This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The Weil Correspondence and
Universal Special Geometry

Abstract

The Weil correspondence states that the datum of a Seiberg-Witten differential is equivalent to an algebraic group extension of the integrable system associated to the Seiberg-Witten geometry. Remarkably this group extension represents quantum consistent couplings for the 𝒩=2{\cal N}=2 QFT if and only if the extension is anti-affine in the algebro-geometric sense. The universal special geometry is the algebraic integrable system whose Lagrangian fibers are the anti-affine extension groups; it is defined over a base ℬ\mathscr{B} parametrized by the Coulomb coordinates and the couplings. On the total space of the universal geometry there is a canonical (holomorphic) Euler differential. The ordinary Seiberg-Witten geometries at fixed couplings are symplectic quotients of the universal one, and the Seiberg-Witten differential arises as the reduction of the Euler one in accordance with the Weil correspondence. This universal viewpoint allows to study geometrically the flavor symmetry of the 𝒩=2{\cal N}=2 SCFT in terms of the Mordell-Weil lattice (with NΓ©ron-Tate height) of the Albanese variety A𝕃A_{\mathbb{L}} of the universal geometry seen as a quasi-Abelian variety Y𝕃Y_{\mathbb{L}} defined over the function field 𝕃≑ℂ​(ℬ)\mathbb{L}\equiv{\mathbb{C}}(\mathscr{B}).

1 Introduction and Overview

In this paper β€œspecial geometry” stands for the holomorphic integrable system111Β A holomorphic integral system is an even-dimensional complex manifold 𝒳\mathscr{X} equipped with a non-degenerate, holomorphic, closed (2,0)(2,0)-form Ξ©\Omega. In our set-up 𝒳\mathscr{X} is in addition algebraic. (𝒳,Ξ©)(\mathscr{X},\Omega) associated to the Seiberg-Witten geometry of a 4d 𝒩=2{\cal N}=2 QFT [SW1, SW2, donagi0, donagi1]. To avoid special cases (which require a slightly different treatment) we make once and for all the assumption that no subsector of the QFT has a weakly coupled Lagrangian formulation.222Β This is a β€œrigidity” condition. It is equivalent to the requirement that the geometry is a deformation of a β„‚Γ—{\mathbb{C}}^{\times}-isoinvariant special geometry with no β„‚Γ—{\mathbb{C}}^{\times}-isoinvariant deformations. The basic message of this paper is that the traditional approach to special geometry does not completely capture the richness of the theory, and fails to address several issues, while there is a more elegant geometric framework where these questions are fully dealt with (at least in principle).

There is an on-going program [A1, A2, A3, A4, A5, A6, A7, A8, A9, A10, A11, A12, A13, A14, A15, A16, A17, A18, A19, A20, A21, A22, A23] to classify all existing 𝒩=2{\cal N}=2 SCFT using special geometry. The holy grail of the program is to construct all 𝒩=2{\cal N}=2 SCFTs with no stringy or Lagrangian construction (or rule out their existence). While partially successful, this geometric program suffers from a serious drawback: to a given scale-invariant special geometry there correspond several inequivalent SCFTs with distinct flavor symmetry groups, central charges, etc. To complete the program one needs a geometric understanding of the set of SCFTs described by a a given scale-invariant special geometry. The traditional viewpoint gives no clue on this issue. There are many other features of SCFTs with no geometric intepretation in that set-up. To make a simple example: from QFT we know that the relevant chiral deformations of a SCFT are in one-to-one correspondence with the elements of the chiral ring of dimension 1<Ξ”<21<\Delta<2. How do we understand this elementary fact in purely geometric terms? We also know that the mass deformations span the Cartan subalgebra of the flavor symmetry group FF. What is the geometric interpretation of these deformations, and how we may use them to determine geometrically the allowed FF’s? In this direction an open problem is to determine the maximal rank fmaxf_{\text{max}} of the flavor group for a rank-rr SCFT. What is fmaxf_{\text{max}} geometrically?

Going to details, there are many aspects of the standard story which don’t look fully satisfactory. One is the Seiberg-Witten differential Ξ»\lambda. The traditional viewpoint is that Ξ»\lambda is a meromorphic differential on 𝒳\mathscr{X} whose periods are the (local) special coordinates. This looks not to be the complete story on two counts:

  • (a)

    for a SCFT Ξ»\lambda is more properly an infinitesimal symmetry of the special geometry 𝒳\mathscr{X}. An infinitesimal symmetry of 𝒳\mathscr{X} is given by a holomorphic vector field β„°βˆˆπ”žβ€‹π”²β€‹π”±β€‹(𝒳){\cal E}\in\mathfrak{aut}(\mathscr{X}). In a symplectic manifold vectors and 1-forms are equivalent, and we can write our infinitesimal symmetry as the 1-form λ≑ιℰ​Ω\lambda\equiv\iota_{\cal E}\Omega, hiding its role as the generator of the connected component of π– π—Žπ—β€‹(𝒳)\mathsf{Aut}(\mathscr{X}).333Β We stress that since Ξ»\lambda is the generator of π– π—Žπ—β€‹(𝒳)0\mathsf{Aut}(\mathscr{X})^{0}, it is canonically determined by the underlying symplectic manifold (𝒳,Ξ©)(\mathscr{X},\Omega) and not an additional datum. Away from conformality the relation of Ξ»\lambda with the geometric symmetries seems to be lost, while one suspects that it must hold in general in the appropriate sense.

  • (b)

    In general one has444Β This equation is meant in the sense of cohomology of currents. Note that the (2,0)(2,0) form Ξ©\Omega represents a (1,1)(1,1) class not a (2,0)(2,0) one.

    [Ξ©]=[d​λ]=βˆ‘ama​Da[\Omega]=[d\lambda]=\sum_{a}m_{a}D_{a} (1.1)

    where the DaD_{a} are polar divisors of Ξ»\lambda and the coefficients (residues) mam_{a} are mass parameters. It has been observed by Ron Donagi [donagi1] that this expression agrees with the Duistermaat-Heckman (DH) formula [DH, cannas]: the cohomology class of the symplectic form depends linearly on the values of the momentum map. Now, while the parameters mam_{a} are, in a broad sense, momentum maps, the DH theorem refers to a particular geometric situation namely the Marsden-Weinstein-Meyer (MWM) symplectic quotient [cannas]. In the usual formulation of special geometry the manifold 𝒳\mathscr{X} is not a MWM quotient of some bigger integrable system, so something looks missing from the picture.

Taking seriously Donagi’s suggestion, one concludes that there should exist a much bigger integrable system in the form of a universal special geometry (𝒴,𝛀)(\mathscr{Y},\boldsymbol{\Omega}) of which the ordinary one (𝒳,Ξ©)(\mathscr{X},\Omega) is a symplectic reduction. The existence of the universal geometry also follows from standard physical considerations, see the paragraph below about the physical interpretation via spurions. In the spirit of item (a) one expects that in the universal geometry the differential 𝝀\boldsymbol{\lambda} (or rather its dual vector field β„°{\cal E}) generates holomorphic automorphisms of 𝒴\mathscr{Y}, so 𝝀\boldsymbol{\lambda} must be canonically determined by the geometry of 𝒴\mathcal{Y}. In particular the polar classes [Da][D_{a}] are uniquely fixed by (𝒴,𝛀)(\mathscr{Y},\boldsymbol{\Omega}). Since we can read the flavor symmetry group FF from the [Da][D_{a}]’s, going to the universal geometry (𝒴,𝛀)(\mathscr{Y},\boldsymbol{\Omega}) should also solve (in principle) the issue of a geometric understanding of flavor symmetry.

The story is a bit more complicated since, in addition to the mass deformations (associated to the flavor symmetry), we have the relevant couplings that also require a geometric interpretation. They should also be determined by the universal symmetry β„°βˆˆπ”žβ€‹π”²β€‹π”±β€‹(𝒴){\cal E}\in\mathfrak{aut}(\mathscr{Y}) of the bigger geometric framework. In other words, there must be an β€œuniversal” geometric theory which encompasses all kinds of physically allowed interactions and singles out the ones which are consistent at the full quantum level.

Our proposal for the solution of these issues (and many others) is based on an old mathematical gadget: the β€œWeil correspondence” that we shall review in a moment after some preliminary.

The fields of definition 𝕂\mathbb{K} and 𝕃\mathbb{L}.

We make the observation that an ordinary special geometry 𝒳\mathscr{X} is, in particular, a complex model of an Abelian variety (or scheme) X𝕂X_{\mathbb{K}} defined over the field 𝕂=ℂ​(u1,…,ur)\mathbb{K}={\mathbb{C}}(u_{1},\dots,u_{r}) of rational functions in rr variables (rr being the rank i.e.​ the dimension of the ordinary Coulomb branch π’ž\mathscr{C}). This assertion contains a mild but crucial assumption that we now make explicit. In all known 𝒩=2{\cal N}=2 SCFT the Coulomb branch555Β Here we look at the Coulomb branch as a mere complex manifold, stripped of all other structures such as metric etc. π’ž\mathscr{C} is a copy of β„‚r{\mathbb{C}}^{r}. However, as pointed out in ref.​[A8], there are reasons to believe that more general Coulomb branches are also allowed. Since the Coulomb branch π’ž\mathscr{C} is the spectrum of the chiral ring β„›\mathscr{R}, in any quantum-consistent 𝒩=2{\cal N}=2 SCFT π’ž\mathscr{C} is an irreducible, reduced, normal, affine (complex) variety with an algebraic β„‚Γ—{\mathbb{C}}^{\times}-action (see section 2 for more details).

Question 1.

Which (possibly singular) affine varieties are the Coulomb branch π’ž\mathscr{C} of some physically sound 𝒩=2{\cal N}=2 QFT?

The allowed π’ž\mathscr{C}’s are severely restricted. If you look to a reasonable-looking candidate Coulomb branch ≄ℂr\not\simeq{\mathbb{C}}^{r} you almost certainly end up with the conclusion that the dimension of the deformation space of the special geometry is not equal to the number of chiral operators of dimension ≀2\leq 2 as required by QFT (and true for π’žβ‰ƒβ„‚r\mathscr{C}\simeq{\mathbb{C}}^{r} [A23]). The β€œexperimental” evidence indicate that in each dimension rr there are few putative Coulomb branches consistent with this constraint from the dimension of the deformation space; all of them are expected to have the same field of functions666Β In particular in rank-11 there is only one possible Coulomb branch: the plane β„‚{\mathbb{C}} [A11]. ℂ​(π’ž)≑ℂ​(u1,…,ur){\mathbb{C}}(\mathscr{C})\equiv{\mathbb{C}}(u_{1},\dots,u_{r}). In this paper we make this empirical suggestion into an assumption. Under this assumption, while the geometry of the Coulomb branch may be rather subtle, the subtleties are confined in codimension at least two,777Β The singularities, if present, are in codimension β‰₯2\geq 2 since we are free to assume π’ž\mathscr{C} to be normal by replacing the chiral ring β„›\mathscr{R} with its normalization, a procedure which is also suggested by QFT considerations. while the aspects we are interested in (at this early stage of the program) are typically codimension-1 phenomena.

Since we are mainly interested in codimension-1 issues, we don’t bother which specific model of the Abelian variety X𝕂X_{\mathbb{K}} yields the actual special geometry 𝒳\mathscr{X}. At this stage we can work with X𝕂X_{\mathbb{K}} which is simply a commutative group (defined on the slightly fancier field 𝕂\mathbb{K}) and get a lot of mileage from algebraic group theory [milne].

However the focus of this paper is not the usual special geometry 𝒳\mathscr{X}, but the much bigger universal integrable system 𝒴\mathscr{Y}. Just as we may replace 𝒳\mathscr{X} with the underlying commutative group X𝕂X_{\mathbb{K}} defined over the function field 𝕂=ℂ​(π’ž)\mathbb{K}={\mathbb{C}}(\mathscr{C}), we may replace 𝒴\mathscr{Y} by a commutative algebraic group Y𝕃Y_{\mathbb{L}} defined over the bigger function field 𝕃=ℂ​(ℬ)\mathbb{L}={\mathbb{C}}(\mathscr{B}) which is an extension of 𝕂\mathbb{K} of transcendental degree equal to the complex dimension of the space of mass and relevant couplings. The group-(scheme) Y𝕃Y_{\mathbb{L}} is our main object of interest.

The Seiberg-Witten differential Ξ»\lambda.

In the algebraic group language the Seiberg-Witten differential Ξ»\lambda is a differential on X𝕂X_{\mathbb{K}} defined over 𝕂\mathbb{K}. From the QFT point of view Ξ»\lambda depends on three distinct kinds of variables:888Β  There are no marginal couplings under our assumption that theory has no Lagrangian subsectors. the Coulomb coordinates (u1,…,ur)(u_{1},\dots,u_{r}), the chiral couplings (t1,…,tk)(t_{1},\dots,t_{k}), and the mass parameters (m1,…,mf)(m_{1},\dots,m_{f}). The geometric distinction between the three kinds of parameters is as follows:

βˆ‚Ξ»βˆ‚ui\displaystyle\frac{\partial\lambda}{\partial u_{i}} ∈{differentials of the first kind on ​X𝕂}\displaystyle\in\Big{\{}\text{differentials of the first kind on }X_{\mathbb{K}}\Big{\}} (1.2)
βˆ‚Ξ»βˆ‚tj\displaystyle\frac{\partial\lambda}{\partial t_{j}} ∈{differentials of the second kind on ​X𝕂}{first kindΒ +Β exact differentials}\displaystyle\in\frac{\Big{\{}\text{differentials of the second kind on }X_{\mathbb{K}}\Big{\}}}{\Big{\{}\text{first kind $+$ exact differentials}\Big{\}}}
βˆ‚Ξ»βˆ‚ma\displaystyle\frac{\partial\lambda}{\partial m_{a}} ∈{differentials of the third kind on ​X𝕂}{second kindΒ +Β exact differentials}\displaystyle\in\frac{\Big{\{}\text{differentials of the third kind on }X_{\mathbb{K}}\Big{\}}}{\Big{\{}\text{second kind $+$ exact differentials}\Big{\}}}

The parameters uiu_{i}, tjt_{j} and mam_{a} exhaust the supply of differentials, so in Ξ»\lambda there is no room for couplings of fancier kinds. This is consistent with the superspace construction of 𝒩=2{\cal N}=2 QFT. We shall see below that, in addition, only chiral couplings consistent with quantum UV-completeness are allowed in special geometry.

Eq.(1.2) suggests that we should treat the parameters uiu_{i}, tjt_{j} and mam_{a} on the same footing. Therefore we introduce the universal branch β„¬β‰ƒπ’žΓ—π’«\mathscr{B}\simeq\mathscr{C}\times\mathscr{P} with global999Β To make the masses mam_{a} into global coordinates we may need to go to a finite cover of the physical coupling space. coordinates uiu_{i}, tjt_{j}, mam_{a}. We write 𝕃=ℂ​(ℬ)≃ℂ​(ui,tj,ma)\mathbb{L}={\mathbb{C}}(\mathscr{B})\simeq{\mathbb{C}}(u_{i},t_{j},m_{a}) for the rational field in r+k+fr+k+f variables. Over ℬ\mathscr{B} we have the universal Abelian variety

\xymatrixβ€‹π’œβ€‹\xy@@ix@β€‹βˆ’\xymatrix{\mathscr{A}\xy@@ix@{{\hbox{}}}}{-} (1.3)
\PATHlabelsextra@@

[r]^(0.3)Ο–&B≃CΓ—P, i.e.​ the family of ordinary special geometries parametrized by the coupling space 𝒫\mathscr{P}. Again, π’œβ†’β„¬\mathscr{A}\to\mathscr{B} is a model of an Abelian variety A𝕃A_{\mathbb{L}} defined over 𝕃\mathbb{L}. The Seiberg-Witten differential Ξ»\lambda is then a third-kind differential on A𝕃A_{\mathbb{L}} defined over 𝕃\mathbb{L}. The ordinary special geometry at fixed couplings p≑(tΒ―j,mΒ―a)βˆˆπ’«p\equiv(\underline{t}_{j},\underline{m}_{a})\in\mathscr{P} is π’³β‰‘π’œ|π’žΓ—p\mathscr{X}\equiv\mathscr{A}|_{\mathscr{C}\times p}.

The Weil correspondence.

There are many math gadgets known under the name β€œWeil correspondence”. The one of interest in this paper is perhaps less known. To put it in the proper perspective, we quote from the book by Mazur and Messing [MM]:101010Β The references in the quote insides the oval box are: Weil paper [Weil]; Barsotti paper [barsotti], Serre paper [serre1] and Serre book [serre2].

In [27] Weil observed that when working on abelian variety AA over an arbitrary field, considerations of extensions of AA by a vector group replaces the study of differentials of the second kind, while considerations of extensions of AA by a torus replaces the study of differentials of the third kind. He attributes these ideas (in the classical case) to Severi. Over β„‚\mathbb{C}, Barsotti in [1 bis] established algebraically the isomorphism 𝖀𝗑𝗍​(A,𝔾a)β‰…differentials of second kindholomorphic differentials+exact differentials\mathsf{Ext}(A,\mathbb{G}_{a})\cong\frac{\text{differentials of second kind}}{\text{holomorphic differentials}+\text{exact differentials}} (See Serre’s [24] and [25] for a beautiful account of these ideas)

Comparing this quotation with eq.(1.2) we conclude that we may identify the couplings allowed by 𝒩=2{\cal N}=2 supersymmetry with commutative algebraic groups. In particular the ordinary Seiberg-Witten geometry (𝒳,Ξ»)(\mathscr{X},\lambda) (with mass and relevant deformations switched on) defines (and is fully determined by) an algebraic group Z𝕂Z_{\mathbb{K}} over 𝕂\mathbb{K}. To physicists group means symmetry, and we think the differential Ξ»\lambda as describing a symmetry of the geometry. Again, to get a more intrinsic and satisfactory picture one has to work with the appropriate commutative extension Y𝕃Y_{\mathbb{L}} of the universal Abelian group-variety A𝕃A_{\mathbb{L}} defined over the field 𝕃\mathbb{L}. The physical meaning of Y𝕃Y_{\mathbb{L}} will be clarified below.

Anti-affine (quasi-Abelian) groups.

The Weil correspondence replaces the issue of a geometric description of the allowed physical couplings (in particular mass deformations) with the more geometric question of understanding the allowed symmetries. However not all commutative algebraic groups Y𝕃Y_{\mathbb{L}} are allowed for the following reasons. Consider the mass deformations of a SCFT. Physically they are associated to the flavor Lie group FF: the mass parameters take value in the Cartan algebra π” β€‹π”žβ€‹π”―β€‹(F)\mathfrak{car}(F) of FF. The deformation space π” β€‹π”žβ€‹π”―β€‹(F)\mathfrak{car}(F) comes with a number of specific structures: an action of the Weyl group, an invariant bilinear form, and an integral lattice. In addition, we expect that the rank of FF is bounded in each fixed rank r≑dimπ’žr\equiv\dim\mathscr{C}. Generic torus extensions do not come with such structures, and their rank has no upper bound. Hence a general extension is not allowed. The case of relevant couplings is even sharper. According to the Weil correspondence, a SCFT with the relevant couplings switched on (but no mass deformation) corresponds to a group Y𝕃Y_{\mathbb{L}} which is a vector extension of the universal Abelian variety A𝕃A_{\mathbb{L}}. An Abelian variety has extensions by vector groups of arbitrary dimension. At the classical level the 𝒩=2{\cal N}=2 SCFT has indeed infinitely many linearly independent chiral deformations, but almost all of them are inconsistent at the quantum level because they spoil UV completeness. The deformations which are consistent at the full non-perturbative level are in one-to-one correspondence with the chiral operators of the undeformed theory with scaling dimension 1<Ξ”<21<\Delta<2, in particular their number is at most rr. This leads to the

Question 2.

Which algebraic groups Y𝕃Y_{\mathbb{L}} over 𝕃\mathbb{L} are the Weil correspondents of 𝒩=2{\cal N}=2 QFTs which are fully consistent at the quantum level (UV complete)?

The fact that all 4d 𝒩=2{\cal N}=2 QFT can be twisted Γ‘ la Witten [ttf1, witten, marino] into a Topological Field Theory (TFT), leads to the following (see main text):

Answer.

The algebraic group Y𝕃Y_{\mathbb{L}} must be anti-affine (a.k.a.​ quasi-Abelian), that is,

Γ​(Y𝕃,π’ͺY𝕃)≃𝕃.\Gamma(Y_{\mathbb{L}},{\cal O}_{Y_{\mathbb{L}}})\simeq\mathbb{L}. (1.4)

Here π’ͺY𝕃{\cal O}_{Y_{\mathbb{L}}} is the structure sheaf of the algebraic variety Y𝕃Y_{\mathbb{L}} defined over 𝕃\mathbb{L}. In this paper we check that (1.4) exactly matches the conditions from quantum consistency of the QFT.

Universal special geometry.

Up to now we just considered (families of) ordinary special geometries, but this is clearly not the full story as the Donagi remark in item (b) indicates. There must be a bigger algebraic integrable system with underlying algebraic group Y𝕃Y_{\mathbb{L}}. The dimension (over 𝕃\mathbb{L}) of Y𝕃Y_{\mathbb{L}} is the total number of parameters r+k+fr+k+f. A model over β„‚{\mathbb{C}} of Y𝕃Y_{\mathbb{L}} will be a fibration

Ο€:π’΄β†’β„¬β‰ƒπ’žΓ—π’«,\pi\colon\mathscr{Y}\to\mathscr{B}\simeq\mathscr{C}\times\mathscr{P}, (1.5)

where the fibers and the base ℬ\mathscr{B} both have complex dimension r+k+fr+k+f. We claim that the total space 𝒴\mathscr{Y} is a β€œsymplectic variety” with symplectic form 𝛀\boldsymbol{\Omega}, while the fibers of Ο€\pi are Lagrangian submanifolds. We write β€œsymplectic variety” between quotes because we do not have control on singularities in codimension β‰₯2\geq 2 and it is possible (even expected) that for some 𝒩=2{\cal N}=2 QFT these singularities have no crepant resolution.111111Β A crepant resolution is automatically a symplectic variety. Let us make the claim more precise. The Lagrangian fibration (1.5) has a β„‚Γ—{\mathbb{C}}^{\times} group of automorphisms. Let β„°{\cal E} be the Euler vector field, that is, the generator of the Lie algebra π”žβ€‹π”²β€‹π”±β€‹(𝒴)\mathfrak{aut}(\mathscr{Y}) normalized so that

ℒℰ​𝛀=𝛀.\mathscr{L}_{\cal E}\boldsymbol{\Omega}=\boldsymbol{\Omega}. (1.6)

The dual 1-form 𝝀=ιℰ​𝛀\boldsymbol{\lambda}=\iota_{\cal E}\boldsymbol{\Omega} is then the universal differential which is holomorphic and canonically defined by the symmetries of the geometry. From (1.6) d​𝝀=𝛀d\boldsymbol{\lambda}=\boldsymbol{\Omega}. The Lagrangian fibration (1.5) is our universal special geometry. The only difference with respect to an ordinary special geometry is that now the generic (smooth) fiber is a general quasi-Abelian variety instead of an Abelian variety.

Symplectic quotients and SW differentials.

The Lagrangian fibration (1.5) is an (algebraic) Liouville integrable system whose smooth fibers may be non-compact. Its Hamiltonians in involution are the regular functions on the affine base ℬ\mathscr{B}. We can perform the symplectic quotient at fixed values tΒ―j,mΒ―a\underline{t}_{j},\underline{m}_{a} of the couplings tjt_{j} and masses mam_{a} setting

π’ž={(ui,tj,ma)βˆˆβ„¬:tj=tΒ―j,ma=mΒ―a}βŠ‚β„¬,\displaystyle\mathscr{C}=\big{\{}(u_{i},t_{j},m_{a})\in\mathscr{B}\colon t_{j}=\underline{t}_{j},m_{a}=\underline{m}_{a}\big{\}}\subset\mathscr{B}, (1.7)
π’³β‰‘Ο€βˆ’1​(π’ž)/Hβ†’π’ž,\displaystyle\mathscr{X}\equiv\pi^{-1}(\mathscr{C})/H\to\mathscr{C}, (1.8)

where HH is the group generated by the vector fields dual to the differentials d​tjdt_{j}’s and d​madm_{a}’s. The reduced fibration π’³β†’π’ž\mathscr{X}\to\mathscr{C} is an ordinary special geometry over the ordinary Coulomb branch π’ž\mathscr{C} whose symplectic form class [Ξ©][\Omega] depends linearly on the parameters tjt_{j}, mam_{a} (in facts only on the masses mam_{a}) according to the Duistermaat-Heckman formula [DH]. More in detail: on 𝒴\mathscr{Y} we have the God given universal Euler differential 𝝀\boldsymbol{\lambda} which induces a differential Ξ»\lambda on the symplectic reduction 𝒳\mathscr{X} such that d​λ=Ξ©d\lambda=\Omega. The differential Ξ»\lambda arising from the symplectic quotient coincides with the usual Seiberg-Witten differential: in this framework the Seiberg-Witten differential is constructed out of the symmetries of the problem. The construction of Ξ»\lambda is very explicit: the basic tool is Picard’s construction of anti-affine groups. It is easy to check in the examples that Ξ»\lambda is the physically correct differential. The reader may be puzzled. The universal differential 𝝀\boldsymbol{\lambda} is perfectly holomorphic. How it happens that Ξ»\lambda has now become meromorphic?. The point is that all group extension is a principal bundle, and to write explicit expressions we need to choose a gauge on this bundle. Just as for the magnetic monopole in ℝ3{\mathbb{R}}^{3}, if you insist to use a single coordinate chart you are forced to pick up a singular gauge, and the gauge connection AA will look singular in that gauge; but the singularity is a mere gauge artifact: in a regular gauge it looks perfectly regular. The Seiberg-Witten differential is singular (in presence of non-trivial couplings) just because we write it in a (convenient) singular gauge.

Physical interpetation: Spurions.

Finally we give the physical motivations for the extension of the special geometry from the ordinary version 𝒳\mathscr{X} to the universal one 𝒴\mathscr{Y}. From the viewpoint of the superspace approach to SUSY QFTs, we can always see the couplings as 𝒩=2{\cal N}=2 chiral superfields which are frozen to their constant vev’s. Such non-dynamical superfields are usually called spurions. In particular the masses may be thought of as complex scalars in SYM superfields which weakly gauge the flavor symmetry FF: physically we may see the flavor symmetry as the zero-coupling limit of a gauge symmetry. In other words, to switch on the mass deformations we gauge the Cartan subgroup121212Β Properly speaking one has to gauge the full group FF; however in the Ξ»β†’0\lambda\to 0 limit the difference becomes inessential. of FF with rank​F\mathrm{rank}\,F spurion vector supermultiplets, adding a kinetic term for them

1gf2β€‹βˆ«(βˆ’14​Fμ​ν​a​Fμ​ν​a+βˆ‚ΞΌMaβˆ—β€‹βˆ‚ΞΌMa+fermions)​d4​x,\frac{1}{g_{f}^{2}}\int\!\Big{(}-\frac{1}{4}F_{\mu\nu\,a}\,F^{\mu\nu\,a}+\partial_{\mu}M_{a}^{*}\,\partial^{\mu}M_{a}+\text{fermions}\Big{)}d^{4}x, (1.9)

and then send gfβ†’0g_{f}\to 0 while keeping fixed ⟨Ma⟩=mΒ―a\langle M_{a}\rangle=\underline{m}_{a}. In the same way the relevant couplings tjt_{j} can be seen as 𝒩=2{\cal N}=2 chiral superfields TjT_{j} which are frozen to constant values by rescaling their kinetic terms131313Β The TiT_{i}’s kinetic terms are irrelevant operators; the situation is rather similar to the case of (2,2) Landau-Ginzburg in two dimensions [ttstar]. by an overall factor 1/Ο΅21/\epsilon^{2} and then sending Ο΅β†’0\epsilon\to 0 while keeping fixed ⟨Tj⟩=tΒ―j\langle T_{j}\rangle=\underline{t}_{j}. The interaction with couplings tjt_{j} is written as an integral over the 𝒩=2{\cal N}=2 chiral superspace

βˆ‘j∫d4​x​d4​θ​Tj​ϕj+h.c.β†’Ο΅β†’0βˆ‘jtjβ€‹βˆ«d4​x​d4​θ​ϕj+h.c.\sum_{j}\int d^{4}x\,d^{4}\theta\,T_{j}\,\phi_{j}+\text{h.c.}\xrightarrow{\ \epsilon\to 0\ }\sum_{j}t_{j}\int d^{4}x\,d^{4}\theta\,\phi_{j}+\text{h.c.} (1.10)

where the Ο•j\phi_{j}’s are the chiral superfields whose first components are the operators ujβˆˆβ„›u_{j}\in\mathscr{R} of dimension 1<Ξ”j<21<\Delta_{j}<2. If we add only vector spurions, at finite gfg_{f} we have just an ordinary141414Β More precisely the germ of an ordinary special geometry since after the gauging of the flavor group the theory is typically non UV-complete. Since this is a physicists’ argument, we dispense the reader with too many technical pedantries. special geometry with a Coulomb branch ℬ\mathscr{B} of dimension r+fr+f. Sending gfβ†’0g_{f}\to 0 has two effects:

  • (1)

    it friezes the Coulomb coordinates associated to the flavor group to the values mΒ―aβ‰‘βŸ¨Ma⟩\underline{m}_{a}\equiv\langle M_{a}\rangle, thus replacing ℬ\mathscr{B} with the subvariety π’ž={ma=mΒ―a}\mathscr{C}=\{m_{a}=\underline{m}_{a}\} i.e.​ with the ordinary Coulomb branch;

  • (2)

    the (generic) fiber over π’ž\mathscr{C} degenerates from an Abelian variety of dimension r+fr+f to a semi-Abelian variety (≑\equiv a torus extension of an Abelian variety). A semi-Abelian variety arises as a semi-stable degeneration of an Abelian variety, so (in a sense) this is the mildest possible degeneration for an ordinary special geometry.

The situation with relevant coupling is similar except that the dimension of the corresponding operator is Ξ”>1\Delta>1 so the fiber degeneration is more severe and we get a vector extension (an unstable degeneration). Before sending gf,Ο΅β†’0g_{f},\epsilon\to 0 both the masses mam_{a} and the chiral couplings tjt_{j} are generators of the universal chiral ring 𝒬\mathscr{Q}. In the limit these generators are replaced by complex numbers, and 𝒬\mathscr{Q} specializes to the usual chiral ring β„›\mathscr{R}. The base ℬ\mathscr{B} of the extended geometry is the affine variety 𝖲𝗉𝖾𝖼​𝒬\mathsf{Spec}\,\mathscr{Q}.

Open problems.

The Weil correspondence solves many issues but it also opens new questions. The most important one is to find the maximal rank f​(r)f(r) of the flavor group FF as a function of the dimension rr of the ordinary Coulomb branch. The overall picture is roughly as follows. The projective closure of (a suitable model of) X𝕂X_{\mathbb{K}} is expected to be a Fano variety β„±{\cal F} (possibly singular) of dimension 2​r2r with the property that β„±βˆ–K{\cal F}\setminus K is symplectic (KK a canonical divisor). E.g.​ for r=1r=1 this condition says that β„±{\cal F} is a rational elliptic surface [A11]. The rank of the flavor group is related to the Picard number of β„±{\cal F}. Hence the question boils down to an optimal upper bound on the Picard numbers of Fano varieties of dimension 2​r2r with β„±βˆ–K{\cal F}\setminus K symplectic and only physically admissible singularities.

Organization of the paper.

The rest of the paper contains detailed explanations, computations, and technicalities making the above picture more precise. In section 2 we review special geometry from our abstract viewpoint. In section 3 we discuss the anti-affine algebraic groups, study Lagrangian-fibrations whose general fibers are anti-affine groups, and describe their physical interpretation. In section 4 we introduce the Mordell-Weil lattices and their NΓ©ron-Tate height pairing. We then compare the resulting structures with the physics of flavor symmetry. In section 5 we compute the Seiberg-Witten differential using the Weil correspondence and the Picard explicit construction of quasi-Abelian groups.

2 General special geometry

General special geometry may be summarized as follows. We have a polarized, normal algebraic variety 𝒴\mathscr{Y} over β„‚{\mathbb{C}} with a holomorphic symplectic 2-form 𝛀\boldsymbol{\Omega}. β€œPolarized” means equipped with an integral class Ο‰βˆˆH2​(𝒴,β„€)∩H1,1​(𝒴)\omega\in H^{2}(\mathscr{Y},{\mathbb{Z}})\cap H^{1,1}(\mathscr{Y}) containing a KΓ€hler metric. The crucial property is that the affinization morphism

Ο€:𝒴→𝖲𝗉𝖾𝖼​Γ​(𝒴,π’ͺ𝒴)≑ℬ\pi\colon\mathscr{Y}\to\mathsf{Spec}\,\Gamma(\mathscr{Y},\mathscr{O}_{\mathscr{Y}})\equiv\mathscr{B} (2.1)

is a fibration with (connected) Lagrangian fibers, hence 𝒴\mathscr{Y} is an algebraic Liouville integrable system whose first integrals of motion in involution are precisely all the regular functions on the β€˜phase space’ 𝒴\mathscr{Y}. The (holomorphic) Hamiltonian vector fields generate an action G↷𝒴G\curvearrowright\mathscr{Y} of a connected commutative group GG and Ο€\pi is its momentum map. Thus

The relevant geometries are (in particular) algebraic integrable systems whose momentum map ΞΌ\mu coincides with the affinization morphism Ο€\pi.

We write 𝒬≑Γ​(𝒴,π’ͺ𝒴)\mathscr{Q}\equiv\Gamma(\mathscr{Y},\mathscr{O}_{\mathscr{Y}}) and call it the universal ring. The fiber 𝒴b\mathscr{Y}_{b} over a generic point bβˆˆβ„¬b\in\mathscr{B} is smooth. The locus of points bβˆˆβ„¬b\in\mathscr{B} with non-smooth fiber, iff non-empty, is a divisor π’ŸβŠ‚β„¬\mathscr{D}\subset\mathscr{B} called the discriminant. When Ο€\pi has a section, e:ℬ→𝒴e\colon\mathscr{B}\to\mathscr{Y}, the smooth locus of a fiber π’΄ΜŠbβŠ‚π’΄b\mathring{\mathscr{Y}}_{b}\subset\mathscr{Y}_{b} is an algebraic group whose connected component has the form G/Ge​(b)G/G_{e(b)}, where Ge​(b)βŠ‚GG_{e(b)}\subset G is the discrete isotropy subgroup. e​(b)e(b) is then the neutral element of the group π’΄ΜŠb0\mathring{\mathscr{Y}}_{b}^{0}, and ee is called the zero-section. We take the existence of a zero-section such that e​(ℬ)βŠ‚π’΄e(\mathscr{B})\subset\mathscr{Y} is Lagrangian as part of our definition of special geometry; for more general situations see [A22]. Then 𝒴\mathscr{Y} may be seen as a polarized commutative group-scheme over the affine scheme ℬ\mathscr{B}. The special geometries of interest in this paper have, in addition, an action ℂ↷𝒴{\mathbb{C}}\curvearrowright\mathscr{Y} by conformal-symplectic automorphisms such that as 𝛀↦ez​𝛀\boldsymbol{\Omega}\mapsto e^{z}\,\boldsymbol{\Omega} (zβˆˆβ„‚z\in{\mathbb{C}}). The quotient group which acts faithfully (and algebraically) is a copy of the multiplicative group ℂ×≃ℂ/2​π​i​m​℀{\mathbb{C}}^{\times}\simeq{\mathbb{C}}/2\pi im{\mathbb{Z}}. The Lie algebra of β„‚Γ—{\mathbb{C}}^{\times} is generated by a Euler vector field β„°{\cal E} such that

ℒℰ​𝛀=𝛀\mathscr{L}_{\cal E}\boldsymbol{\Omega}=\boldsymbol{\Omega} (2.2)

to which it corresponds a differential 𝝀=ιℰ​𝛀\boldsymbol{\lambda}=\iota_{\cal E}\boldsymbol{\Omega} with d​𝝀=𝛀d\boldsymbol{\lambda}=\boldsymbol{\Omega}. The smooth fibers 𝒴b\mathscr{Y}_{b} are connected commutative group-varieties over β„‚{\mathbb{C}}. The universal cover of all such group-varieties is β„‚n{\mathbb{C}}^{n}. Then, analytically, all smooth fibers 𝒴b\mathscr{Y}_{b} can be written as β„‚n/Ξ›b{\mathbb{C}}^{n}/\Lambda_{b} for some discrete subgroup Ξ›bβŠ‚β„‚n\Lambda_{b}\subset{\mathbb{C}}^{n}. If (x1,β‹―,xn)(x^{1},\cdots,x^{n}) are affine coordinates in the covering β„‚n{\mathbb{C}}^{n}, the Euler differential 𝝀\boldsymbol{\lambda} locally takes the Darboux form

𝝀=βˆ‘ipi​d​xi⇒𝛀=d​pi∧d​xi\boldsymbol{\lambda}=\sum_{i}p_{i}\,dx^{i}\quad\Rightarrow\quad\boldsymbol{\Omega}=dp_{i}\wedge dx^{i} (2.3)

for suitable local functions pip_{i} in ℬ\mathscr{B} with ℒℰ​pi=pi\mathscr{L}_{\cal E}\mspace{1.0mu}p_{i}=p_{i}. We take the above statements as our basic assumptions (β€œaxioms”).

Physical justifications.

All the statements above may be justified (or at least argued) from the very first principles of quantum physics. In particular the equality of the momentum map and the affinization map, eq.(2.1), follows from the fact that an 𝒩=2{\cal N}=2 QFT can be twisted into a Topological Field Theory (TFT) Γ‘ la Witten [ttf1, witten, marino]. After the twist, the TFT has the following properties:

  • (1)

    the IR effective theory is exact for topological amplitudes;

  • (2)

    the TFT amplitudes on ℝ3Γ—S1{\mathbb{R}}^{3}\times S^{1} are independent of the radius RR of the circle. As Rβ†’βˆžR\to\infty we can use the 4d IR effective theory which is the usual Seiberg-Witten description with Coulomb branch ℬ\mathscr{B} [wittenM]. In the limit Rβ†’0R\to 0 we can use the 3d IR effective theory which is a Οƒ\sigma-model with the hyperKΓ€hler target (𝒴,𝛀)(\mathscr{Y},\boldsymbol{\Omega}) [gaiotto2]. The results of the two computations should agree, in particular we must have the same ring of local topological observables 𝒬\mathscr{Q} in both descriptions. This is eq.(2.1).

The basic idea is that we may twist the 𝒩=2{\cal N}=2 theory obtained by promoting the couplings to spurion superfields. For instance, we can see the masses as the result of very weak gauging of the flavor symmetry and topologically twist this weakly gauged theory. As the gauge coupling goes to zero the Abelian fibers of the integrable system degenerate into semi-Abelian ones in the well-known way. Abstract special geometry (as defined above) captures all sectors of the 𝒩=2{\cal N}=2 model except for free hypermultiplets which are decoupled from the rest of the theory and hence do not talk with the vector multiplets which parametrize the special geometry. This paper is dedicated to understanding the structure of the geometric objects satisfying the β€œaxioms”. We start by a review of the ordinary Seiberg-Witten geometries.

2.1 Review: the case of proper fibers

In all abstract special geometries a generic fiber 𝒴b\mathscr{Y}_{b} is smooth, hence a complex algebraic group with the property that it has no non-constant regular functions (cf.​ (2.1)). The obvious way to get rid of all non-constant regular functions is to take the fiber to be compact (proper). In this case the generic fiber 𝒴b\mathscr{Y}_{b} is an Abelian variety with polarization Ο‰|𝒴b\omega|_{\mathscr{Y}_{b}}. In most applications one assumes the polarization to be principal, but the story makes sense (geometrically as well as physically) for polarizations of any degree. When the generic fiber is proper we use the standard notations and terminology: we write β„›\mathscr{R} (resp.​ π’ž\mathscr{C}) for 𝒬\mathscr{Q} (resp.​ for ℬ\mathscr{B}) which we call the chiral ring (resp.​ Coulomb branch); the total space of the geometry will be written 𝒳\mathscr{X} with symplectic form Ξ©\Omega. The dimension rr of π’ž\mathscr{C} is the rank of the geometry.

2.1.1 Superconformal geometries

As already mentioned, we are particularly interested in special geometries whose automorphism group contains β„‚Γ—{\mathbb{C}}^{\times}. They describe 𝒩=2{\cal N}=2 SCFTs where all mass and relevant deformations are switched off. The Lie algebra of the automorphism group β„‚Γ—{\mathbb{C}}^{\times} is generated by a complete holomorphic vector field β„°{\cal E} (the Euler vector) such that

ℒℰ​Ω=Ξ©\mathscr{L}_{\cal E}\mspace{1.0mu}\Omega=\Omega (2.4)

which implies

Ξ©=d​λℰwhereλℰ≑ιℰ​Ω.\Omega=d\lambda_{\cal E}\quad\text{where}\quad\lambda_{\cal E}\equiv\iota_{\cal E}\Omega. (2.5)

The Euler differential Ξ»β„°\lambda_{\cal E} is the Seiberg-Witten (SW) differential for SCFT geometries. The β„‚Γ—{\mathbb{C}}^{\times}-action on the regular functions induces a β„‚Γ—{\mathbb{C}}^{\times}-action on π’ž\mathscr{C}. A Hamiltonian hβˆˆβ„›h\in\mathscr{R} has dimension Δ​(h)\Delta(h) iff ℒℰ​h=Δ​(h)​h\mathscr{L}_{\cal E}h=\Delta(h)\,h. The identity has dimension 0, all other elements of hβˆˆβ„›h\in\mathscr{R} must have dimension

Δ​(h)β‰₯1\Delta(h)\geq 1 (2.6)

this fact is known as the unitary bound (see below). β„›\mathscr{R} is then a graded ring of the form

β„›=β„‚β‹…1βŠ•β„›+,\displaystyle\mathscr{R}={\mathbb{C}}\cdot 1\oplus\mathscr{R}_{+}, β„›+=⨁1β‰€Ξ”βˆˆ1n​ℕℛΔ,\displaystyle\mathscr{R}_{+}=\bigoplus_{1\leq\Delta\in\frac{1}{n}\mathbb{N}}\mathscr{R}_{\Delta}, (2.7)
β„›Ξ”β‹…β„›Ξ”β€²βŠ‚β„›Ξ”+Ξ”β€²\displaystyle\mathscr{R}_{\Delta}\cdot\mathscr{R}_{\Delta^{\prime}}\subset\mathscr{R}_{\Delta+\Delta^{\prime}} hβˆˆβ„›Ξ”β‡”β„’β„°β€‹h=Δ​h,\displaystyle h\in\mathscr{R}_{\Delta}\ \Leftrightarrow\ \mathscr{L}_{\cal E}\,h=\Delta\,h,

where we used that nβ€‹Ξ”βˆˆβ„•n\,\Delta\in\mathbb{N} for some integer nn since β„‚Γ—{\mathbb{C}}^{\times} acts algebraically. It is known that only finitely many nn may appear for a given rr [A9]. The maximal ideal β„›+βŠ‚β„›\mathscr{R}_{+}\subset\mathscr{R} corresponds to the closed point 0βˆˆπ’ž0\in\mathscr{C}, called the origin, which is the only closed β„‚Γ—{\mathbb{C}}^{\times}-orbit in π’ž\mathscr{C}. If we require π’ž\mathscr{C} to be smooth at 0, we get that ℛ≃ℂ​[u1,…,ur]\mathscr{R}\simeq{\mathbb{C}}[u_{1},\dots,u_{r}] is a free polynomial ring, hence π’ž\mathscr{C} is β„‚r{\mathbb{C}}^{r} with coordinates uiu_{i} of dimension Ξ”iβ‰₯1\Delta_{i}\geq 1. However π’ž\mathscr{C} may be singular (cf. Question 1). We write βˆ‚Ο†i\partial_{\varphi^{i}} for the Hamiltonian vector vuiv_{u_{i}}. The dual differentials d​φid\varphi^{i} have dimension

Δ​(d​φi)=βˆ’Ξ”β€‹(vui)=1βˆ’Ξ”i.\Delta(d\varphi_{i})=-\Delta(v_{u_{i}})=1-\Delta_{i}. (2.8)

Let 𝒳u\mathscr{X}_{u} be a smooth fiber. Consider the exponential map

expe​(u)(ziβˆ‚Ο†i):𝖫𝗂𝖾(𝒳u)≑Te​(u)𝒳uβŸΆπ’³u,uβˆˆπ’ž.\exp_{e(u)}\mspace{-5.0mu}\Big{(}z^{i}\partial_{\varphi^{i}}\Big{)}\colon\mathsf{Lie}(\mathscr{X}_{u})\equiv T_{e(u)}\mathscr{X}_{u}\longrightarrow\mathscr{X}_{u},\qquad u\in\mathscr{C}. (2.9)

and let KuK_{u} be its kernel. Clearly

𝒳u≃𝖫𝗂𝖾​(𝒳u)/Ku,that is,Ο†iβˆΌΟ†i+A​(u)i​j​(mj+τ​(u)j​k​nk),mj,nkβˆˆβ„€,\mathscr{X}_{u}\simeq\mathsf{Lie}(\mathscr{X}_{u})/K_{u},\quad\text{that is,}\quad\varphi^{i}\sim\varphi^{i}+A(u)^{ij}(m_{j}+\tau(u)_{jk}\,n^{k}),\quad m^{j},n^{k}\in{\mathbb{Z}}, (2.10)

where τ​(u)i​j\tau(u)_{ij} is the period matrix of the Abelian fiber at uu, and

ℒℰ​τi​j=0\mathscr{L}_{\cal E}\mspace{2.0mu}\tau_{ij}=0 (2.11)

since β„°{\cal E} acts by automorphisms. The formula (2.10) holds also when 𝒳u\mathscr{X}_{u} is not smooth, except that 𝖫𝗂𝖾​(𝒳u)/Ku\mathsf{Lie}(\mathscr{X}_{u})/K_{u} should be identified with the connected component of the smooth locus of the fiber containing e​(u)e(u). Now,

ℒℰ​Ai​j=(1βˆ’Ξ”i)​Ai​jfor all ​j.\mathscr{L}_{\cal E}\,A^{ij}=(1-\Delta_{i})A^{ij}\quad\text{for all }j. (2.12)
Unitary bound.

Let us show eq.(2.6). All complete vector vv satisfies ℒℰ​v=Δ​(v)​v\mathscr{L}_{\cal E}v=\Delta(v)\,v with Δ​(v)β‰₯0\Delta(v)\geq 0, while for the Hamiltonian vector vhv_{h} such that ΞΉvh​Ω=d​hβ‰’0\iota_{v_{h}}\Omega=dh\not\equiv 0 one has

0≀Δ​(vh)≑Δ​(h)βˆ’1.0\leq\Delta(v_{h})\equiv\Delta(h)-1. (2.13)

The elements uiβˆˆβ„›u_{i}\in\mathscr{R} which saturate the unitary bound (2.6), i.e.​ such that Ξ”i=1\Delta_{i}=1, have special properties. From eq.(2.12) we see that Ξ”i=1\Delta_{i}=1 implies that Ai​jA^{ij} is constant in the closure of any β„‚Γ—{\mathbb{C}}^{\times}-orbit, hence constant in π’ž\mathscr{C} because 0 belongs to the closure of all orbits. Since the connected component of the smooth locus of the fiber over 0 must be a group, it follows that the fiber over 0 contains an Abelian variety of dimension equal to the multiplicity of 11 as a Coulomb dimension, which then is a constant Abelian subvariety BB contained in all fibers. This Abelian subvariety, describes a free sub-sector which we may decouple without loss. Conversely a constant Abelian subvariety BB contained in all the fibers describes a free sector and dimB=dimβ„›1\dim B=\dim\mathscr{R}_{1}. Indeed let AΞ±A_{\alpha}, BΞ²B^{\beta} be a symplectic basis of H1​(A0,β„€)H_{1}(A_{0},{\mathbb{Z}}). The periods

aΞ±=∫Aαιℰ​Ω,aΞ²D=∫Bβιℰ​Ωa^{\alpha}=\int_{A_{\alpha}}\iota_{\cal E}\Omega,\qquad a^{D}_{\beta}=\int_{B^{\beta}}\iota_{\cal E}\Omega (2.14)

are global regular functions on π’ž\mathscr{C} of dimension 1.151515Β Notice that the free subsector has no non-trivial mass or relevant deformations. We conclude that we may assume with no loss that the chiral ring has the form ℂ​[u1,…,ur]{\mathbb{C}}[u_{1},\dots,u_{r}], where rβ‰₯1r\geq 1 and Ξ”i>1\Delta_{i}>1 for all ii. In this situation there is a non-zero divisor π’ŸβŠ‚π’ž\mathscr{D}\subset\mathscr{C}, called the discriminant, consisting of the points uβˆˆπ’žu\in\mathscr{C} whose fiber 𝒳u\mathscr{X}_{u} is not smooth. The complement π’žΜŠβ‰‘π’žβˆ–π’Ÿ\mathring{\mathscr{C}}\equiv\mathscr{C}\setminus\mathscr{D} (the β€œgood” locus) parametrizes a family of Abelian varieties, hence over π’žΜŠ\mathring{\mathscr{C}} we have a local system with fiber H1​(𝒳u,β„€)H^{1}(\mathscr{X}_{u},{\mathbb{Z}}). The polarization induces a non-degenerate skew-symmetric pairing of this local system

βŸ¨βˆ’,βˆ’βŸ©:H1​(𝒳u,β„€)Γ—H1​(𝒳u,β„€)β†’β„€,⟨ξ,η⟩=βˆ’βŸ¨Ξ·,ξ⟩,\langle-,-\rangle\colon H_{1}(\mathscr{X}_{u},{\mathbb{Z}})\times H_{1}(\mathscr{X}_{u},{\mathbb{Z}})\to{\mathbb{Z}},\qquad\langle\xi,\eta\rangle=-\langle\eta,\xi\rangle, (2.15)

which is physically interpreted as the Dirac electro-magnetic pairing. The fiber H1​(𝒳u,β„€)H_{1}(\mathscr{X}_{u},{\mathbb{Z}}) is the lattice of quantized electro-magnetic charges in the SUSY preserving vacuum uβˆˆπ’žΜŠu\in\mathring{\mathscr{C}}. The map

Zu:H1​(𝒳u,β„€)β†’β„‚,Zu​(Ξ±)=βˆ«Ξ±ΞΉβ„°β€‹Ξ©,Z_{u}\colon H_{1}(\mathscr{X}_{u},{\mathbb{Z}})\to{\mathbb{C}},\qquad Z_{u}(\alpha)=\int_{\alpha}\iota_{\cal E}\Omega, (2.16)

is the SUSY central charge of a state with electro-magnetic charge Ξ±\alpha.

2.1.2 The Chow 𝕂/β„‚\mathbb{K}/{\mathbb{C}}-trace

For later reference we rephrase the saturation of the unitary bound in terms of the unerlying Abelian variety X𝕂X_{\mathbb{K}} defined over the function field 𝕂\mathbb{K}. An invariant of an Abelian variety X𝕂X_{\mathbb{K}} defined over a complex function field 𝕂≑ℂ​(π’ž)\mathbb{K}\equiv{\mathbb{C}}(\mathscr{C}) is its Chow 𝕂/β„‚\mathbb{K}/{\mathbb{C}}-trace [trace1, trace2, langII, langIII]

tr𝕂/ℂ​(X𝕂)≑(Bβ„‚,Ο„)\mathrm{tr}_{\mathbb{K}/{\mathbb{C}}}(X_{\mathbb{K}})\equiv(B_{\mathbb{C}},\tau) (2.17)

which is an Abelian variety Bβ„‚B_{\mathbb{C}} defined over β„‚{\mathbb{C}} together with a map defined over 𝕂\mathbb{K}

Ο„:Bβ„‚β†’X𝕂\tau\colon B_{\mathbb{C}}\to X_{\mathbb{K}} (2.18)

which satisfies the appropriate universal mapping property. Here we are interested in the physical meaning of the Chow trace (see also [A11, A23]). We claim that the Chow trace of X𝕂X_{\mathbb{K}} is the Abelian variety Bβ„‚B_{\mathbb{C}} of the free subsector. Indeed under the isomorphism

Ξ©:𝖫𝗂𝖾​(𝒳u)β†’Tuβˆ—β€‹π’žβ‰ƒd​ℛ/d​ℛ2,\Omega\colon\mathsf{Lie}(\mathscr{X}_{u})\to T_{u}^{*}\mathscr{C}\simeq d\mathscr{R}/d\mathscr{R}^{2}, (2.19)

which sends Hamiltonian vector fields to the differentials of their Hamiltonians, the image of the Lie algebra of the Chow trace Ω​(𝖫𝗂𝖾​(Bβ„‚))\Omega(\mathsf{Lie}(B_{\mathbb{C}})) is contained in d​ℛ1d\mathscr{R}_{1} since the Hamiltonian vectors βˆ‚w\partial_{w} tangent to the Chow-trace have dimension zero. Dually all Hamiltonians in β„›1\mathscr{R}_{1} generate constant vertical vector field. Thus

Fact 1.

In absence of free subsectors

tr𝕂/ℂ​(X𝕂)=0.\mathrm{tr}_{\mathbb{K}/{\mathbb{C}}}(X_{\mathbb{K}})=0. (2.20)

2.1.3 Switching on mass and relevant deformations

After switching on the mass and relevant deformations the geometry 𝒳\mathscr{X} remains a holomorphic Lagrangian fibration with section over the (same) Coulomb branch π’ž\mathscr{C} with generic Abelian fibers. However the β„‚Γ—{\mathbb{C}}^{\times} symmetry is no longer present, and the Euler holomorphic differential λℰ≑ιℰ​Ω\lambda_{\cal E}\equiv\iota_{\cal E}\Omega gets replaced by a meromorphic Seiberg-Witten differential Ξ»SW\lambda_{\text{SW}} which depends on the Coulomb branch coordinates uiu_{i}, the masses mam_{a}, and the relevant couplings tjt_{j}. Eq.(2.16) still holds with the replacement of Ξ»β„°\lambda_{\cal E} by Ξ»SW\lambda_{\text{SW}}. Ξ»SW\lambda_{\text{SW}} is defined modulo exact forms and its dependence on the various parameters follows the rule (1.2). This rule shows that the tangent space to π’ž\mathscr{C} (resp.​ to the space of relevant couplings) injects in H0​(𝒳u,Ξ©1)H^{0}(\mathscr{X}_{u},\Omega^{1}) (resp.​ H1​(𝒳u,π’ͺ)H^{1}(\mathscr{X}_{u},\mathcal{O})). In particular (as it is clear from the physical side) the space of relevant deformations of a SCFT has dimension at most rr and is uniquely determined by the β„‚Γ—{\mathbb{C}}^{\times}-invariant special geometry of the SCFT. The story with the mass deformations is much more involved. This reflects the physical fact that the mass deformation of a SCFT geometry is non-unique, in general, that is, there are several distinct flavor symmetries (of different ranks) which are consistent with one and the same SCFT special geometry. This non-uniqueness can be seen per tabulas in the explicit classification of the flavor symmetries of rank-1 SCFTs in refs.[A3, A4, A5, A6, A7, A8, A11, A13]. This concludes our quick review of the ordinary special geometries with compact generic fibers (for more see e.g.​ [A23]). Next we consider the general case where the generic fiber may be non-compact.

3 Geometries with anti-affine fibers

The ordinary geometries described above are not the only ones which satisfy our physically motivated β€œaxioms”. The basic condition on the algebraic integrable system that the Liouville momentum map is the affinization morphism Ο€\pi has other solutions with group-variety fibers. One of the goals of this paper is to present the physical interpretation of these more general geometries. We first describe them geometrically, beginning with the structure of a single generic fiber.

3.1 Anti-affine groups

An algebraic group is anti-affine [milne] iff it has no non-constant regular functions; anti-affine groups are also known as quasi-Abelian varieties. An anti-affine group is automatically smooth, connected, and commutative [milne]. Our basic β€œaxiom”, eq.(2.1), says that a general smooth fiber 𝒴b\mathscr{Y}_{b} is an anti-affine group. This is consistent with the interpretation of the geometry as a complex integrable system: the Lie algebra of the fiber is commutative being generated by Hamiltonians in involution. However the fibers of most algebraic integrable systems are not anti-affine.

Remark 1.

In the literature there are two distinct notions of β€œquasi-Abelian variety”. A complex variety is quasi-Abelian in the algebraic sense iff it has no non-constant regular function. It is quasi-Abelian in the analytic sense iff its underlying complex manifold has no non-constant holomorphic function. Clearly the analytic notion is stronger, and (as we shall see below) there exist group-varieties which are quasi-Abelian in the algebraic sense but have non-trivial holomorphic functions, in facts whose underlying complex manifold is Stein. We stress that the notion relevant for our applications is the algebraic one. The existence of non-trivial holomorphic functions which are not algebraic does not spoil the quantum consistency of the QFT associated with the geometry.

To describe the structure of an anti-affine group over β„‚{\mathbb{C}} we start from the structure of a general connected, commutative, complex algebraic group GG which may or may not be anti-affine. The Barsotti-Chevalley theorem [milne] states that GG is an extension of group-varieties of the form

eβ†’Cβ†’Gβ†’π–Ίπ—…π– β†’πŸ’,e\to C\xrightarrow{\phantom{mm}}G\xrightarrow{\sf\;\,al\;}A\to 0, (3.1)

where AA is an Abelian variety, in facts the Albanese variety of GG (𝖺𝗅\mathsf{al} is the Albanese map [langI, milne2]), and CC is a connected commutative affine group with identity ee. A connected affine commutative group CC is a product of copies of additive and multiplicative groups 𝔾ak×𝔾mf\mathbb{G}_{a}^{k}\times\mathbb{G}_{m}^{f}; when working over β„‚{\mathbb{C}} we identify it with the complex Lie group

β„‚kΓ—(β„‚Γ—)f≑VΓ—T,whereV=β„‚k,T=(β„‚Γ—)f.{\mathbb{C}}^{k}\times({\mathbb{C}}^{\times})^{f}\equiv V\times T,\quad\text{where}\ \ V={\mathbb{C}}^{k},\quad T=({\mathbb{C}}^{\times})^{f}. (3.2)

We call T≑𝔾mfT\equiv\mathbb{G}_{m}^{f} the torus group and V≑𝔾akV\equiv\mathbb{G}_{a}^{k} the vector group. The inequivalent commutative group-varieties GG with given AA, kk and ff are then classified by the group

𝖀𝗑𝗍1​(A,VΓ—T)=𝖀𝗑𝗍1​(A,V)βŠ•π–€π—‘π—1​(A,T).\mathsf{Ext}^{1}(A,V\times T)=\mathsf{Ext}^{1}(A,V)\oplus\mathsf{Ext}^{1}(A,T). (3.3)

The extension groups are computed in [serre2]

𝖀𝗑𝗍1​(A,T)\displaystyle\mathsf{Ext}^{1}(A,T) ≃𝖯𝗂𝖼0​(A)f≑(A∨)f\displaystyle\simeq\mathsf{Pic}^{0}(A)^{f}\equiv(A^{\vee})^{f} (3.4)
𝖀𝗑𝗍1​(A,V)\displaystyle\mathsf{Ext}^{1}(A,V) ≃VβŠ•dimA,\displaystyle\simeq V^{\oplus\dim A}, (3.5)

where A∨A^{\vee} is the dual Abelian variety. For later reference we sketch the proof of (3.4),(3.5).

Sketch of proof of (3.4),(3.5).

Recall that 𝖯𝗂𝖼0​(A)\mathsf{Pic}^{0}(A) is the group of divisors on AA which are algebraically equivalent to zero modulo linear equivalence, while Aβˆ¨β‰…π–―π—‚π–Ό0​(A)A^{\vee}\cong\mathsf{Pic}^{0}(A) is the dual Abelian variety of AA (see e.g.​ Β§. I.8 of [milne2]). One shows161616Β See Proposition 8.4 of [milne2], or Β§. 5.2 of [langII]. that an invertible sheaf (≑\equiv line bundle ≑\equiv linear class of divisors) β„’βˆˆπ–―π—‚π–Όβ€‹(A){\cal L}\in\mathsf{Pic}(A) is algebraically trivial if and only if it is invariant under translations in AA, that is, if for all a∈A​(β„‚)a\in A({\mathbb{C}}) we have

taβˆ—β€‹β„’β‰ƒβ„’,where ​ta:b↦b+a,a,b∈A​(β„‚).t_{a}^{*}{\cal L}\simeq{\cal L},\quad\text{where }t_{a}\colon b\mapsto b+a,\ \ a,b\in A({\mathbb{C}}). (3.6)

We first consider extensions of AA by 𝔾m\mathbb{G}_{m}. Complex analytically, GG is a principal bundle over AA with structure group β„‚Γ—{\mathbb{C}}^{\times}, which we may see as a bundle β„’βˆ—β†’A{\cal L}^{*}\to A where β„’βˆ—{\cal L}^{*} is the complement of the zero section in a line bundle β„’β†’A{\cal L}\to A. The total space of a principal bundle over an Abelian group AA with Abelian structure group β„‚Γ—{\mathbb{C}}^{\times} is itself an Abelian group if and only if the line bundle is invariant by translation on AA so that the two group operations along the base and the fiber commute; by the remark before eq.(3.6) this is equivalent to β„’{\cal L} being algebraically trivial. Next we consider extensions of AA by 𝔾a\mathbb{G}_{a}. Complex analytically GG is a principal bundle over AA with group β„‚{\mathbb{C}}, and all such principal bundles are commutative groups. Hence the extension group is the space of isomorphism classes of such bundles H1​(A,π’ͺA)≃ℂdimAH^{1}(A,{\cal O}_{A})\simeq{\mathbb{C}}^{\dim A}. ∎

We note that

H1​(G,β„€)≃℀2​dimA+dimT.H_{1}(G,{\mathbb{Z}})\simeq{\mathbb{Z}}^{2\dim A+\dim T}. (3.7)

The algebraic group GG in eq.(3.1) is connected and commutative but may or may not be anti-affine. For instance if AA is trivial, GG is affine; more generally when the extension is trivial, G=CΓ—AG=C\times A, the group is not anti-affine, etc. Our next task is to understand which extensions are anti-affine. We consider first the extension of AA by the ff-torus T≑(β„‚Γ—)fT\equiv({\mathbb{C}}^{\times})^{f} and then the extension by a vector group V≃ℂkV\simeq{\mathbb{C}}^{k}.

3.1.1 Torus extensions of Abelian varieties

If GG is an extension of the Abelian variety AA by the torus group TT

eβ†’Tβ†’Gβ†’π–Ίπ—…π– β†’πŸ’e\to T\to G\xrightarrow{\sf\;al\;}A\to 0 (3.8)

the Albanese map G→𝖺𝗅𝖠G\xrightarrow{\sf\;al\;}A is a principal TT-bundle. We ask when a TT-principal bundle over a complex Abelian variety AA is a commutative algebraic group. Let Ξ›\Lambda be the character group of T≃(β„‚Γ—)fT\simeq({\mathbb{C}}^{\times})^{f}. For any TT-principal bundle G→𝛼AG\xrightarrow{\,\alpha\,}A we have the β€œFourier decomposition”

Ξ±βˆ—β€‹π’ͺG=β¨Ξ»βˆˆΞ›β„’Ξ»,\alpha_{\ast}\mspace{1.0mu}{\cal O}_{G}=\bigoplus_{\lambda\in\Lambda}{\cal L}_{\lambda}, (3.9)

where β„’Ξ»β†’A{\cal L}_{\lambda}\to A are line bundles (invertible sheaves). Around eq.(3.6) we saw that the translations in AA are TT-automorphisms of the principal bundle GG iff the line bundles β„’Ξ»β†’A{\cal L}_{\lambda}\to A are algebraically equivalent to zero, that is, β„’Ξ»βˆˆπ–―π—‚π–Ό0​(A){\cal L}_{\lambda}\in\mathsf{Pic}^{0}(A) for all Ξ»βˆˆΞ›\lambda\in\Lambda. Stated differently:

Lemma 1 (See e.g.​ [brion]).

The TT-principal bundle G→𝛼AG\xrightarrow{\;\alpha\;}A is a group homomorphism with kernel TT if and only if the decomposition (3.9) defines a group homomorphism

Ο‡:Λ→𝖯𝗂𝖼0​(A),Ο‡:λ↦ℒλ.\chi\colon\Lambda\to\mathsf{Pic}^{0}(A),\qquad\chi\colon\lambda\mapsto{\cal L}_{\lambda}. (3.10)

Then we have

Γ​(G,π’ͺG)=H0​(A,Ξ±βˆ—β€‹π’ͺG)=β¨Ξ»βˆˆΞ›H0​(A,β„’Ξ»)\Gamma(G,{\cal O}_{G})=H^{0}(A,\alpha_{\ast}{\cal O}_{G})=\bigoplus_{\lambda\in\Lambda}H^{0}(A,{\cal L}_{\lambda}) (3.11)

while

H0​(A,β„’0)≑H0​(A,π’ͺA)=β„‚.H^{0}(A,{\cal L}_{0})\equiv H^{0}(A,{\cal O}_{A})={\mathbb{C}}. (3.12)

Hence GG is anti-affine iff H0​(A,β„’Ξ»)=0H^{0}(A,{\cal L}_{\lambda})=0 for all Ξ»β‰ 0\lambda\neq 0. Since H0​(A,β„’)=0H^{0}(A,{\cal L})=0 for all line bundle β„’{\cal L} which is algebraically trivial but non-trivial, we need that β„’Ξ»{\cal L}_{\lambda} is the zero element of 𝖯𝗂𝖼0​(A)\mathsf{Pic}^{0}(A) only when Ξ»=0\lambda=0, that is,

Lemma 2.

The group GG in eq.(3.8) is anti-affine if and only if the underlying group homomorphism Ο‡:Λ→𝖯𝗂𝖼0​(A)\chi\colon\Lambda\to\mathsf{Pic}^{0}(A) is injective, i.e.​ if and only if Ξ›\Lambda is a lattice in the dual Abelian variety A∨A^{\vee}.

An example will clarify the situation.

Example 3.1 (Picard 1910).

We consider the extensions of the elliptic curve EΟ„E_{\tau} of period Ο„\tau by the one-dimensional torus β„‚Γ—{\mathbb{C}}^{\times}. We see EΟ„E_{\tau} as the quotient of its universal cover, β„‚{\mathbb{C}}, by the group generated by the two automorphisms

L1:z↦z+1andL2:x↦z+Ο„,L_{1}\colon z\mapsto z+1\quad\text{and}\quad L_{2}\colon x\mapsto z+\tau, (3.13)

and we write Ο€:β„‚β†’Eτ≑ℂ/⟨L1,L2⟩\pi\colon{\mathbb{C}}\to E_{\tau}\equiv{\mathbb{C}}/\langle L_{1},L_{2}\rangle for the canonical projection. Let p1:β„‚Γ—β„‚β†’β„‚p_{1}\colon{\mathbb{C}}\times{\mathbb{C}}\to{\mathbb{C}} be the trivial β„‚{\mathbb{C}}-bundle. A line bundle β„’wβ†’EΟ„{\cal L}_{w}\to E_{\tau} which is algebraically trivial can always be written as the quotient

β„’w=(β„‚Γ—β„‚)/⟨L1,L2⟩,L1:(z,y)↦(z+1,y)L2:(z,y)β†’(z+Ο„,e2​π​i​w​y).{\cal L}_{w}=({\mathbb{C}}\times{\mathbb{C}})\big{/}\langle L_{1},L_{2}\rangle,\qquad\begin{aligned} L_{1}&\colon(z,y)\mapsto(z+1,y)\\ L_{2}&\colon(z,y)\to(z+\tau,e^{2\pi iw}y).\end{aligned} (3.14)

From this expression it is obvious that the inequivalent bundles β„’wβˆˆπ–―π—‚π–Ό0​(EΟ„){\cal L}_{w}\in\mathsf{Pic}^{0}(E_{\tau}) are parametrized by the points w∈EΟ„w\in E_{\tau} (which coincides with EΟ„βˆ¨β‰ƒπ–―π—‚π–Ό0​(EΟ„)E_{\tau}^{\vee}\simeq\mathsf{Pic}^{0}(E_{\tau})). To get the corresponding β„‚Γ—{\mathbb{C}}^{\times} extension of EΟ„E_{\tau} we just cut out the zero section of the bundle

Gw=(β„‚Γ—β„‚Γ—)/⟨L1,L2⟩,L1:(z,y)↦(z+1,y)L2:(z,y)β†’(z+Ο„,e2​π​i​w​y).G_{w}=({\mathbb{C}}\times{\mathbb{C}}^{\times})\big{/}\langle L_{1},L_{2}\rangle,\qquad\begin{aligned} L_{1}&\colon(z,y)\mapsto(z+1,y)\\ L_{2}&\colon(z,y)\to(z+\tau,e^{2\pi iw}y).\end{aligned} (3.15)

It may be convenient to write the principal β„‚Γ—{\mathbb{C}}^{\times}-bundle Gwβ†’EΟ„G_{w}\to E_{\tau} in an alternative way which we dub the β€œsingular gauge” ; we parametrize

(z,y)=(z,ϑ​(zβˆ’w,Ο„)ϑ​(z,Ο„)​x)βˆˆβ„‚Γ—(β„™1βˆ–{0,∞})(z,y)=\left(z,\frac{\vartheta(z-w,\tau)}{\vartheta(z,\tau)}x\right)\in{\mathbb{C}}\times\big{(}\mathbb{P}^{1}\setminus\{0,\infty\}\big{)} (3.16)

where171717Β The ΞΈ\theta-function in the rhs is defined as in chapter 20 of DLMF [dlmf]. By construction ϑ​(z,Ο„)\vartheta(z,\tau) has a single zero at the lattice points z=m+n​τz=m+n\tau.

ϑ​(z,Ο„)​=def​θ3​(π​zβˆ’12β€‹Ο€βˆ’12​π​τ|Ο„)\vartheta(z,\tau)\overset{\rm def}{=}\theta_{3}\big{(}\pi z-\tfrac{1}{2}\pi-\tfrac{1}{2}\pi\tau\,\big{|}\,\tau\big{)} (3.17)

and then take the quotient by the action of the group ⟨L~1,L~2⟩\langle\tilde{L}_{1},\tilde{L}_{2}\rangle acting as

L~1:(z,x)↦(z+1,x),L~2:(z,x)↦(z+Ο„,x),\tilde{L}_{1}\colon(z,x)\mapsto(z+1,x),\qquad\tilde{L}_{2}\colon(z,x)\mapsto(z+\tau,x), (3.18)

where xβˆˆβ„‚Γ—x\in{\mathbb{C}}^{\times} is now a global (but singular) coordinate along the fiber. It is clear that the groups extensions GwG_{w} are parametrized by degree-zero divisors of the form

(ϑ​(zβˆ’w,Ο„)ϑ​(z,Ο„))=wβˆ’0,\left(\frac{\vartheta(z-w,\tau)}{\vartheta(z,\tau)}\right)=w-0, (3.19)

i.e.​ by points of EΟ„βˆ¨β‰…EΟ„E_{\tau}^{\vee}\cong E_{\tau}. We conclude:

Corollary 1.

Modulo isomorphism we have one connected, commutative, algebraic group (over β„‚{\mathbb{C}}), GwG_{w}, which fits in the exact sequence

1→ℂ×→Gw→Eτ→01\to{\mathbb{C}}^{\times}\to G_{w}\to E_{\tau}\to 0 (3.20)

per point w∈Eτ≅𝖯𝗂𝖼0​(EΟ„)w\in E_{\tau}\cong\mathsf{Pic}^{0}(E_{\tau}).

The Corollary is just eq.(3.4) with f=1f=1. Now let us see for which points w∈EΟ„w\in E_{\tau} the commutative algebraic group GwG_{w} is anti-affine. Suppose that the point w∈EΟ„w\in E_{\tau} which defines the group GwG_{w} is torsion, that is, m​w=0mw=0 in EΟ„E_{\tau} for some mβˆˆβ„€m\in{\mathbb{Z}}, i.e.

m​w=a+b​τ​in ​ℂ​withΒ a,bβˆˆβ„€.m\,w=a+b\,\tau\ \text{in }\ {\mathbb{C}}\ \text{with $a,b\in{\mathbb{Z}}$.} (3.21)

The function on β„‚Γ—β„‚Γ—{\mathbb{C}}\times{\mathbb{C}}^{\times}

f​(z,y)≑eβˆ’2​π​i​b​z​ymf(z,y)\equiv e^{-2\pi ibz}\,y^{m} (3.22)

is invariant under L1L_{1}, L2L_{2} so f​(z,y)f(z,y) is a non-constant global holomorphic function and GwG_{w} is not anti-affine. However, when ww is not torsion GwG_{w} is manifestly anti-affine.181818Β It is also quasi-Abelian in the analytic sense. This conclusion is in agreement with Lemma 2 (see [brion] for more details). The dual of the Lie algebra of the 2-dimensional group GwG_{w} is generated by two holomorphic differentials d​zdz and d​y/2​π​i​ydy/2\pi iy. H1​(Gw,β„€)≃℀3H_{1}(G_{w},{\mathbb{Z}})\simeq{\mathbb{Z}}^{3} is generated by the three cycles

A\displaystyle A ={(z​(s),y​(s))=(s,1)},\displaystyle=\{(z(s),y(s))=(s,1)\}, B={(z​(s),y​(s))=(τ​s,e2​π​i​s​w)},\displaystyle B=\{(z(s),y(s))=(\tau s,e^{2\pi isw})\}, (3.23)
C\displaystyle C ={(z​(s),y​(s))=(0,e2​π​i​s)},\displaystyle=\{(z(s),y(s))=(0,e^{2\pi is})\}, where​ 0≀s≀1\displaystyle\text{where}\ \ 0\leq s\leq 1

with period matrix

∫A𝑑z\displaystyle\int_{A}dz =1,\displaystyle=1, ∫B𝑑z\displaystyle\int_{B}dz =Ο„,\displaystyle=\tau, ∫C𝑑z\displaystyle\int_{C}dz =0,\displaystyle=0, (3.24)
∫Ad​y2​π​i​y\displaystyle\int_{A}\frac{dy}{2\pi iy} =0,\displaystyle=0, ∫Bd​y2​π​i​y\displaystyle\int_{B}\frac{dy}{2\pi iy} =w,\displaystyle=w, ∫Cd​y2​π​i​y\displaystyle\int_{C}\frac{dy}{2\pi iy} =1,\displaystyle=1,

and the groups GwG_{w} are distinguished by their period of d​y/2​π​i​ydy/2\pi iy on the BB-cycle. GwG_{w} is anti-affine (quasi-Abelian) when this period cannot be written as a+b​τa+b\tau for a,bβˆˆβ„ša,b\in\mathbb{Q}, i.e.​ when the three non-zero periods (3.24) are linearly independent over β„š\mathbb{Q}.

Torus extensions of Abelian varieties are also called semi-Abelian varieties. They may be seen as semistable degenerations of Abelian varieites.

3.1.2 Vector extensions of Abelian varieties

For AA an Abelian variety over β„‚{\mathbb{C}} we have

𝖀𝗑𝗍1​(A,β„‚)≃𝖫𝗂𝖾​(A),\mathsf{Ext}^{1}\mspace{-2.0mu}(A,{\mathbb{C}})\simeq\mathsf{Lie}(A), (3.25)

so that an extension of AA by a vector space VV of dimension kk is specified by a kk-tuple of elements of the Lie algebra of AA. More precisely

Proposition 1 (see [brion] Proposition 2.3).

All extension GG of an Abelian variety AA by a vector group VV fits in a unique commutative diagram

(3.26)

where E​(A)E(A) is the universal vector extension of AA [ros58, MM] (which is an anti-affine group). Then GG is anti-affine if and only if Ξ³\gamma is surjective. The transpose map

Ξ³t:Vβˆ¨β†’H1​(A,π’ͺA)\gamma^{t}\colon V^{\vee}\to H^{1}(A,{\cal O}_{A}) (3.27)

is then injective and we conclude that the anti-affine groups over AA obtained as vector extensions are classified by subspaces of the β„‚{\mathbb{C}}-space

H1​(A,π’ͺA)≃H0​(A,Ξ©A1)βˆ¨β‰ƒ(Teβˆ—β€‹A)βˆ¨β‰‘Te​A≃𝖫𝗂𝖾​(A)H^{1}(A,{\cal O}_{A})\simeq H^{0}(A,\Omega^{1}_{A})^{\vee}\simeq(T^{*}_{e}A)^{\vee}\equiv T_{e}A\simeq\mathsf{Lie}(A) (3.28)

where the first isomorphism is given by the polarization, the second one by translation invariance in the Abelian group AA.

This statement is crucial for the physical interpretation of vector extensions in the context of 𝒩=2{\cal N}=2 QFT. We have non-trivial group extensions of an Abelian variety AA by any vector space β„‚k{\mathbb{C}}^{k} of arbitrary large dimension kk but if we insist that the extension group should be anti-affine, we conclude that kk cannot be larger than r≑dimAr\equiv\dim A.

Remark 2.

The vector extensions are quasi-Abelian varieties in the algebraic sense but not in the analytic sense. In other words, while they have no non-constant regular functions they have plenty of non-constant holomorphic functions, indeed the underlying complex manifold of the universal vector extension is Stein see the Example below.

Example 3.2.

We consider the additive version of the Picard construction for the group

\xymatrix​0​\xy@@ix@β€‹βˆ’\xymatrix{0\xy@@ix@{{\hbox{}}}}{-} (3.29)
\PATHlabelsextra@@

[r] & C\xy@@ix@

-\PATHlabelsextra@@[r] & H_Ξ·\xy@@ix@-\xy@@ix@!Ch\xy@@ix@!Ch>[r] A\xy@@ix@-\xy@@ix@!Ch\xy@@ix@!Ch>