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

Loop Grassmannian of quivers and Compactified Coulomb branch of quiver gauge theory with no framing

Zhijie Dong
(February 2015)
Abstract

Mirković introduced the notion of loop Grassmannian for symmetric integer matrix κ\kappa. It is a two-step limit of the local projective space ZκαZ_{\kappa}^{\alpha}, which generalizes the usual Zastava for a simply laced group GG. The usual loop Grassmannian of GG is recovered when the matrix κ\kappa is the Cartan matrix of GG. On the other hand, Braverman, Finkelberg, and Nakajima showed that the Compactified Coulomb branch 𝐌Qα\mathbf{M}_{Q}^{\alpha} for the quiver gauge theory with no framing also generalizes the usual Zastava. We show that in the case when κ\kappa is the associated matrix of the quiver QQ, these two generalizations of Zastava coincide, i.e 𝐌QαZκ(Q)α\mathbf{M}_{Q}^{\alpha}\cong Z_{\kappa(Q)}^{\alpha}.

1 Introduction

1.1 Generalization of loop Grassmannian

Let GG be a reductive group over a field k=k¯k=\overline{k} with char(kk)=0=0. Let G𝒦G_{\mathcal{K}} and G𝒪G_{\mathcal{O}} be its formal loop group and formal arc group, respectively. Define 𝒢(G)=G𝒦/G𝒪\mathcal{G}(G)=G_{\mathcal{K}}/G_{\mathcal{O}} as the loop Grassmannian of GG, which is an ind-scheme over kk.

We consider extensions of the concept of the loop Grassmannian to an arbitrary Kac-Moody group GKMG_{KM}. It is not clear how to define the quotient (GKM)𝒦/(GKM)𝒪(G_{KM})_{\mathcal{K}}/(G_{KM})_{\mathcal{O}} as an ind-scheme. The standard approach is to construct a system of finite dimensional schemes. In [BF10, BFN19], normal slices to certain orbits in the (undefined) loop Grassmannian were considered. In two approaches considered here ([BFN19, Mir23]) these schemes are projective and one constructs the whole loop Grassmannian 𝒢(GKM)\mathcal{G}(G_{KM}) as a certain colimit [Mir23].

1.2 Zastava spaces [FM99]

We recall Zastava spaces ZGZ_{G} for a semisimple simply-connected group GG. In 1.3 and 1.4 we will consider two constructions of generalization of Zastava for a quiver.

1.2.1 Intersections of semiinfinite orbits in 𝒢(G)\mathcal{G}(G)

First, let’s get a feeling how the procedure in 1.1 can be done when GG is a simply-connected group.

We start by fixing a Cartan subgroup TT and a pair of opposite Borel subgroups (B+,B)(B^{+},B^{-}) such that B+B=TB^{+}\cap B^{-}=T. Let N±N^{\pm} be the unipotent subgroups such that B±=TN±.B^{\pm}=TN^{\pm}. For any cocharacter α\alpha of TT, let tαG𝒦t^{\alpha}\in G_{\mathcal{K}} be the point corresponding to α\alpha and tα¯\overline{t^{\alpha}} be the corresponding point in 𝒢(G)\mathcal{G}(G). For two cocharacter λ,μ\lambda,\mu of TT, let Sλ+S^{+}_{\lambda} and SμS^{-}_{\mu} be the N𝒦+N_{\mathcal{K}}^{+}-orbit of the point tλ¯\overline{t^{\lambda}} and the N𝒦N_{\mathcal{K}}^{-}-orbit of the point tμ¯\overline{t^{\mu}}, respectively. Furthermore, let Sλ+¯\overline{S^{+}_{\lambda}} and Sμ¯\overline{S^{-}_{\mu}} be their closures in 𝒢(G).\mathcal{G}(G).

For each cocharacter α\alpha, we define Fα=Sα¯S0+¯F^{\alpha}=\overline{S^{-}_{-\alpha}}\cap\overline{S^{+}_{0}}. Now for cocharacters α,β\alpha,\beta such that αβ\alpha\leq\beta, we have a closed embedding FαFβF^{\alpha}\xhookrightarrow{}F^{\beta}, which we refer to as the growth structure.

Taking the inductive limit of FαF^{\alpha} for α0\alpha\geq 0, we recover S0¯=defS0+¯\overline{S_{0}}\stackrel{{\scriptstyle def}}{{=}}\overline{S^{+}_{0}}. Let S0¯×tαS0¯\overline{S_{0}}\xrightarrow[]{\times t^{\alpha}}\overline{S_{0}} be the multiplication map that maps Sα¯S0¯\overline{S_{-\alpha}}\subset\overline{S_{0}} to S0¯.\overline{S_{0}}. Subsequently, taking the direct limit of S0¯×tαS0¯\overline{S_{0}}\xrightarrow[]{\times t^{\alpha}}\overline{S_{0}} for α0\alpha\geq 0 (the direct system will be referred to as the shift structure), we obtain the entire 𝒢(G)\mathcal{G}(G) when GG is simply connected.

This is a case of the prescription from 1.1 with finite-dimensional schemes FαF^{\alpha} and the ind-system given by the growth structure between FαF^{\alpha} and the shift structure between S0¯.\overline{S_{0}}.

1.2.2 Ordered Zastava ZGZ^{\prime}_{G} from global loop Grassmannian

Consider the global version of the loop Grassmannian 𝒢(G).\mathcal{G}(G). This is the Beilinson-Drinfeld Grassmannian 𝒢C(G)\mathcal{G}^{C}(G) defined for a reductive group GG and a smooth curve CC. It has a natural projection 𝒢C(G)𝜋nCn\mathcal{G}^{C}(G)\xrightarrow[]{\pi}\sqcup_{n\in\mathbb{N}}C^{n}. This space has the so-called factorization property. E.g., 𝒢C(G)a𝒢(G)\mathcal{G}^{C}(G)_{a}\cong\mathcal{G}(G) and 𝒢C(G)(a,b)𝒢C(G)a×𝒢C(G)b\mathcal{G}^{C}(G)_{(a,b)}\cong\mathcal{G}^{C}(G)_{a}\times\mathcal{G}^{C}(G)_{b} for ab.a\neq b. For cocharacters λ1,,λn\lambda_{1},\cdots,\lambda_{n}, we can also define the global version of N𝒦±N^{\pm}_{\mathcal{K}}-orbit111 First, one defines the section of map 𝒢C(G)𝜋Cn\mathcal{G}^{C}(G)\xrightarrow[]{\pi}C^{n} corresponding to λ1,,λn\lambda_{1},\cdots,\lambda_{n}[Zhu09, 1.1.8]. Then one defines the global loop group of NN[AR, 2.2.3]. Sλ1,λ2,,λnBD,±𝒢C(G)S_{\lambda_{1},\lambda_{2},\cdots,\lambda_{n}}^{BD,\pm}\subset\mathcal{G}^{C}(G), their orbit closures Sλ1,λ2,,λnBD,±¯\overline{S_{\lambda_{1},\lambda_{2},\cdots,\lambda_{n}}^{BD,\pm}} and intersections FBDα1,αn=Sα1,,αnBD,¯S0,,0BD,+¯F_{BD}^{\alpha_{1}\cdots,\alpha_{n}}=\overline{S_{-\alpha_{1},\cdots,-\alpha_{n}}^{BD,-}}\cap\overline{S_{0,\cdots,0}^{BD,+}} for cocharacters α1,,αn\alpha_{1},\cdots,\alpha_{n}. E.g., the space Sλ1,λ2BDS_{\lambda_{1},\lambda_{2}}^{BD} is over C2C^{2} and its fiber over (a,b)(a,b) is Sλ1×Sλ2S_{\lambda_{1}}\times S_{\lambda_{2}} for aba\neq b and its fiber over (a,a)(a,a) is Sλ1+λ2S_{\lambda_{1}+\lambda_{2}}222The space Sλ1,λ2BD,+¯\overline{S_{\lambda_{1},\lambda_{2}}^{BD,+}} is over C2C^{2} and its fiber over (a,b)(a,b) is Sλ1¯×Sλ2¯\overline{S_{\lambda_{1}}}\times\overline{S_{\lambda_{2}}} for aba\neq b. But here, we only know that the reduced scheme of its fiber over (a,a)(a,a) is Sλ1+λ2¯\overline{S_{\lambda_{1}+\lambda_{2}}} .. For α=αk1+αkn\alpha=\alpha_{k_{1}}+\cdots\alpha_{k_{n}}, where αki\alpha_{k_{i}} is simple coroot, let nin_{i} be the number of jsj^{\prime}s such that αkj=αi\alpha_{k_{j}}=\alpha_{i}. Let Cα=iICniC^{\alpha}=\prod_{i\in I}C^{n_{i}}. Define the ordered Zastava ZGα=FBDαk1,,αknZ_{G}^{\prime\alpha}=F_{BD}^{\alpha_{k_{1}},\cdots,\alpha_{k_{n}}} over CαC^{\alpha}. Here, the space FBDαk1,,αknF_{BD}^{\alpha_{k_{1}},\cdots,\alpha_{k_{n}}} is canonically isomorphic to FBDαk1,,αknF_{BD}^{\alpha^{\prime}_{k_{1}},\cdots,\alpha^{\prime}_{k_{n}}} for a different order of decomposition α=αk1++αkn\alpha=\alpha^{\prime}_{k_{1}}+\cdots+\alpha^{\prime}_{k_{n}} (they are different as subspaces of 𝒢C(G)\mathcal{G}^{C}(G), through). Hence the ordered Zastava ZGαZ_{G}^{\prime\alpha} is independent of the order, which justifies the notation.

1.2.3 Zastava ZGZ_{G} from quasimaps

It turns out that ZGαZ_{G}^{\prime\alpha} and its limit constructions can be defined without using the Beilinson-Drinfeld Grassmannian 𝒢C(G)\mathcal{G}^{C}(G).

First, the affine Zastava333We will fix a curve CC so we drop CC from this notation. The affine Zastava was called Zastava in [FM99] and Zastava was called compactified Zastava in [BFN19]. ZGaff,αZ^{\text{aff},\alpha}_{G} is the space of based quasimaps from the curve CC to the flag variety of GG of degree α[I]\alpha\in\mathbb{N}[I]. It has a natural map π\pi to C×Iα\mathcal{H}^{\alpha}_{C\times I}, the Hilbert scheme of point of C×IC\times I of length α\alpha. Then Zastava ZGαZ^{\alpha}_{G} is defined as certain compactification (fiberwise with respect to π\pi) of ZGaff,αZ^{\text{aff},\alpha}_{G} ( see remark 3.7 in [BFN19] and the reference therein). Now ZGαZ_{G}^{\alpha} is the partially symmetrized version444The map ZGα𝜋iICniZ_{G}^{\prime\alpha}\xrightarrow[]{\pi}\prod_{i\in I}C^{n_{i}} descends to ZGα=ZGα/iISni𝜋iICni/SniC×Iα.Z_{G}^{\alpha}=Z_{G}^{\prime\alpha}/\prod_{i\in I}S^{n_{i}}\xrightarrow[]{\pi}\prod_{i\in I}C^{n_{i}}/S^{n_{i}}\cong\mathcal{H}^{\alpha}_{C\times I}. Define the intersection Faff,BDα1,αn=Sα1,,αnBD,S0,,0BD,+¯F_{\text{aff},BD}^{\alpha_{1}\cdots,\alpha_{n}}=S_{\alpha_{1},\cdots,\alpha_{n}}^{BD,-}\cap\overline{S_{0,\cdots,0}^{BD,+}}. We have ZGaff,α=Faff,BDαk1,,αknZ_{G}^{\prime\text{aff},\alpha}=F_{\text{aff},BD}^{\alpha_{k_{1}},\cdots,\alpha_{k_{n}}}. Then ZGaff,αZ_{G}^{\text{aff},\alpha} is the partially symmetrized version of ZGaff,αZ_{G}^{\prime\text{aff},\alpha}. of ZGα.Z_{G}^{\prime\alpha}. Construction of affine Zastava ZGaffZ_{G}^{\text{aff}} by quasimap was extended to GG being Kac-Moody group in [BFG06], and we expect that the construction of Zastava extends easily.

In this paper, we will consider two other approaches that recover ZGαZ^{\alpha}_{G} when GG is simply-laced555It is an open question to identify these two approaches with the quasimap construction for any quiver besides ADE quivers. We expect that this identification can be reduced to the identification of three constructions of affine Zastava once the quasimap definition of Zastava has been made. For identification of the Coulomb branch affine Zastava (see 1.4) and quasimap affine Zastava, it suffices to show that the quasimap affine Zastava QmapQaff,α\text{Qmap}^{\text{aff},\alpha}_{Q} is normal or flat over C×I\mathcal{H}_{C\times I}[BFN19]. This is known when QQ is ADE or affine type A [BF14] where they constructed a resolution of QmapQaff,α\text{Qmap}^{\text{aff},\alpha}_{Q} and prove it is normal and Cohen-Macaulay, hence flat over C×I\mathcal{H}_{C\times I}.. These two do not directly deal with GKMG_{KM}, but with a quiver QQ. In the approach in 1.4 below, the growth and shift structure can also be constructed. Hence, we accomplished the prescription in 1.1 and have a definition of 𝒢(GKM)\mathcal{G}(G_{KM}) when GKMG_{KM} is associated with a quiver QQ.

1.3 Compactified Coulomb branch for quiver gauge theory with no framing

This is introduced in [BFN18, BFN19]. For any quiver QQ (possibly with loops) and dimension vector α[I]\alpha\in\mathbb{Z}[I], they defined a convolution algebra on certain equivariant homology space of a certain ind-scheme \mathcal{R}, and showed it is commutative. They consider certain positive part +\mathcal{R}^{+} of \mathcal{R}. Denote the Spec of the convolution subalgebra corresponding to +\mathcal{R}^{+} by MQα.\textsf{M}^{\alpha}_{Q}. They also defined a filtration on this subalgebra and defined 𝐌Qα\mathbf{M}^{\alpha}_{Q} as the Proj of its Rees algebra. When the underlying diagram of the quiver QQ is the Dynkin diagram of GG, they showed that MQα\textsf{M}^{\alpha}_{Q} ( or 𝐌Qα\mathbf{M}^{\alpha}_{Q}) coincides with ZGaff,αZ_{G}^{\text{aff},\alpha} ( or ZGαZ_{G}^{\alpha} resp) for C=𝔸1C=\mathbb{A}^{1} (remark 3.7 in [BFN19]). However, defining the growth structure in this approach requires some work [MW24]. This approach, when accompanied by framing, extends the concept of generalized transversal slices to any quiver QQ [BFN19].

1.4 Local projective spaces [MYZ21, Mir23]

1.4.1 Motivations

We motivate this construction by explaining how to reconstruct ZGαZ^{\alpha}_{G} directly from the Cartan matrix of GG. We abbreviate C×Iα\mathcal{H}_{C\times I}^{\alpha} by α\mathcal{H}^{\alpha}. Now we assume GG is simple and simply connected. The factorizable part of the Picard group of 𝒢C(G)\mathcal{G}^{C}(G) is .\mathbb{Z}. Let 𝒪(1)\mathcal{O}(1) be the positive generator of it. Denote the descent of 𝒪(1)|ZG\mathcal{O}(1)|_{Z^{\prime}_{G}} (under the partial symmetrization map ZGZGZ^{\prime}_{G}\xrightarrow[]{}Z_{G}) to ZGZ_{G} also by 𝒪(1).\mathcal{O}(1). For a sheaf \mathcal{F} on XX and a map XYX\xrightarrow[]{}Y, denote by X/Y\mathcal{F}_{X/Y} the pushforward of \mathcal{F} from XX to YY (the notation is used when the map XYX\xrightarrow[]{}Y is clear from the context).

We have the Kodaira embedding ZGα𝑖((𝒪(1)ZGα/α))Z^{\alpha}_{G}\xrightarrow[]{i}\mathbb{P}((\mathcal{O}(1)_{Z^{\alpha}_{G}/\mathcal{H}^{\alpha}})^{*}). Now we further assume that GG is simply laced. A crucial observation is that the restriction map 𝒪(1)ZGα/α𝒪(1)(ZGα)T/α\mathcal{O}(1)_{Z^{\alpha}_{G}/\mathcal{H}^{\alpha}}\cong\mathcal{O}(1)_{(Z^{\alpha}_{G})^{T}/\mathcal{H}^{\alpha}} is an isomorphism. Moreover, (ZGα)T(Z_{G}^{\alpha})^{T} has a moduli description that solely depends on α\alpha. It is Gr(𝒯α)Gr(\mathcal{T}^{\alpha}), the Hilbert scheme of the tautological bundle 𝒯α\mathcal{T}^{\alpha} over α\mathcal{H}^{\alpha}. The generic fiber of the map ZGα𝜋αZ^{\alpha}_{G}\xrightarrow[]{\pi}\mathcal{H^{\alpha}} is an nn-th power of 1\mathbb{P}^{1}’s, where nn is the length of α\alpha. These 1\mathbb{P}^{1}’s can be regarded as the fibers of Zαik𝜋αikZ^{\alpha_{i_{k}}}\xrightarrow[]{\pi}\mathcal{H}^{\alpha_{i_{k}}}, where α\alpha decomposes into the sum of simple roots αik,1kn,ikI\alpha_{i_{k}},1\leq k\leq n,i_{k}\in I.

Over the regular part of CαC^{\alpha}666This is defined as the pullback under CαC×IαC^{\alpha}\xrightarrow[]{}\mathcal{H}_{C\times I}^{\alpha} of the regular part regα\mathcal{H}_{reg}^{\alpha} of C×Iα\mathcal{H}_{C\times I}^{\alpha}, the line bundle 𝒪(1)\mathcal{O}(1) on ZαZ^{\prime\alpha} is canonically isomorphic to the outer tensor product (over α\mathcal{H}^{\alpha}) of 𝒪(1)\mathcal{O}(1) on these Zαik,1knZ^{\alpha_{i_{k}}},1\leq k\leq n, regardless of GG. Denote the isomorphism by locloc. Consider the ordered Zastava ZGα=ZGα×αCαZ^{\prime\alpha}_{G}=Z^{\alpha}_{G}\times_{\mathcal{H}^{\alpha}}C^{\alpha} (see footnote 4 on page 3). Denote the quotient map CαqααC^{\alpha}\xrightarrow[]{q^{\alpha}}\mathcal{H}^{\alpha}. We have

ZG,regα{Z_{G,reg}^{\prime\alpha}}1××1{\mathbb{P}^{1}\times\cdots\times\mathbb{P}^{1}}((𝒪(1)ZG,regα/Cregα)){\mathbb{P}((\mathcal{O}(1)_{Z^{\prime\alpha}_{G,reg}/C_{reg}^{\alpha}})^{*})}((𝒪(1)𝒪(1)1××1/Cregα)){\mathbb{P}((\mathcal{O}(1)\boxtimes\cdots\boxtimes\mathcal{O}(1)_{\mathbb{P}^{1}\times\cdots\times\mathbb{P}^{1}/C_{reg}^{\alpha}})^{*})}((𝒪(1)(qα)Gr(𝒯regα)/Cregα)){\mathbb{P}((\mathcal{O}(1)_{(q^{\alpha})^{*}Gr(\mathcal{T}_{reg}^{\alpha})/C_{reg}^{\alpha}})^{*})}((𝒪(1)𝒪(1)(pr1)Gr(𝒯αk1)××(prn)Gr(𝒯αkn)/Cregα)){\mathbb{P}((\mathcal{O}(1)\boxtimes\cdots\boxtimes\mathcal{O}(1)_{(pr_{1})^{*}Gr(\mathcal{T}^{\alpha_{k_{1}}})\times\cdots\times(pr_{n})^{*}Gr(\mathcal{T}^{\alpha_{k_{n}}})/C^{\alpha}_{reg}})^{*})}((𝒪(1)(qα)Gr(𝒯α)/Cα)){\mathbb{P}((\mathcal{O}(1)_{(q^{\alpha})^{*}Gr(\mathcal{T}^{\alpha})/C^{\alpha}})^{*})}\scriptstyle{\cong}\scriptstyle{\cong}\scriptstyle{\cong}loc\scriptstyle{loc}\scriptstyle{\cong}loc\scriptstyle{loc}\scriptstyle{\cong}

Here 1\mathbb{P}^{1} means Cregα1\mathbb{P}^{1}_{C_{reg}^{\alpha}} and 1××1\mathbb{P}^{1}\times\cdots\times\mathbb{P}^{1} means the fiber product over CregαC_{reg}^{\alpha}. The space (pri)Gr(𝒯αki)(pr_{i})^{*}Gr(\mathcal{T}^{\alpha_{k_{i}}}) is the pullback of Gr(𝒯αki)Gr(\mathcal{T}^{\alpha_{k_{i}}}) under the projection CregαpriCαki=αki.C_{reg}^{\alpha}\xrightarrow[]{pr_{i}}C^{\alpha_{k_{i}}}=\mathcal{H}^{\alpha_{k_{i}}}. In the left column, the first map is the Kodaira embedding. The second is induced by the isomorphism 𝒪(1)ZGα/Cα𝒪(1)(ZGα)T/Cα𝒪(1)(qα)Gr(𝒯α)/Cα\mathcal{O}(1)_{Z^{\prime\alpha}_{G}/C^{\alpha}}\cong\mathcal{O}(1)_{(Z^{\prime\alpha}_{G})^{T}/C^{\alpha}}\cong\mathcal{O}(1)_{(q^{\alpha})^{*}Gr(\mathcal{T}^{\alpha})/C^{\alpha}}. The third is embedding from the regular part to the whole. In the right column, the first map is the Kodaira embedding. The second map is induced by 𝒪(1)ZGα/Cα𝒪(1)(ZGα)T/Cα𝒪(1)(qα)Gr(𝒯α)/Cα\mathcal{O}(1)_{Z^{\prime\alpha}_{G}/C^{\alpha}}\cong\mathcal{O}(1)_{(Z^{\prime\alpha}_{G})^{T}/C^{\alpha}}\cong\mathcal{O}(1)_{(q^{\alpha})^{*}Gr(\mathcal{T}^{\alpha})/C^{\alpha}} for α=αki\alpha=\alpha_{k_{i}}.

The map in the second row is denoted by locloc which comes from a certain isomorphism (still denoted by) locloc between line bundles. Same for the map locloc on the next row. The isomorphism in the first row is induced by the isomorphism locloc between line bundles that appear in the second row777This isomorphism does not descend to an isomorphism ZG,regαregα1×regα×regαregα1Z_{G,reg}^{\alpha}\cong\mathbb{P}^{1}_{\mathcal{H}_{reg}^{\alpha}}\times_{\mathcal{H}_{reg}^{\alpha}}\cdots\times_{\mathcal{H}_{reg}^{\alpha}}\mathbb{P}^{1}_{\mathcal{H}_{reg}^{\alpha}}.

Under the isomorphism locloc, we get certain sections ScanS_{can} of 𝒪(1)(qα)Gr(𝒯regα)/Cregα\mathcal{O}(1)_{(q^{\alpha})^{*}Gr(\mathcal{T}_{reg}^{\alpha})/C_{reg}^{\alpha}} corresponding to the canonical sections of the outer tensor of 𝒪(1)\mathcal{O}(1) on 1.\mathbb{P}^{1}.

Now it suffices to know how ZregαZ^{\prime\alpha}_{reg} embeds into ((𝒪(1)(qα)Gr(𝒯α)/Cα))\mathbb{P}((\mathcal{O}(1)_{(q^{\alpha})^{*}Gr(\mathcal{T}^{\alpha})/C^{\alpha}})^{*}), since then we can reconstruct ZαZ^{\prime\alpha} as the closure of ZregαZ^{\prime\alpha}_{reg} in ((𝒪(1)(qα)Gr(𝒯α)/Cα))\mathbb{P}((\mathcal{O}(1)_{(q^{\alpha})^{*}Gr(\mathcal{T}^{\alpha})/C^{\alpha}})^{*}) by the fact that the map Zα𝜋CαZ^{\prime\alpha}\xrightarrow[]{\pi}C^{\alpha} is flat. This will be determined by the singularities of these sections ScanS_{can} of 𝒪(1)(qα)Gr(𝒯regα)/Cregα\mathcal{O}(1)_{(q^{\alpha})^{*}Gr(\mathcal{T}_{reg}^{\alpha})/C_{reg}^{\alpha}} along the diagonal of CαC^{\alpha}. The singularities can be classified by a symmetric integer matrix κ\kappa, which is the Cartan matrix of GG.

1.4.2 Construction of Zastava ZκZ_{\kappa} for a matrix κ\kappa

On the other hand, for any symmetric matrix κMI()\kappa\in M_{I}(\mathbb{Z}), there is a line bundle LκL^{\kappa} (2.6) on C×I\mathcal{H}_{C\times I} with certain locality property. Let pr1,pr2pr_{1},pr_{2} be the maps C×Ipr1Gr(𝒯α)pr2C×I\mathcal{H}_{C\times I}\xleftarrow{pr_{1}}Gr(\mathcal{T}^{\alpha})\xrightarrow[]{pr_{2}}\mathcal{H}_{C\times I} such that pr1(DD)=Dpr_{1}(D^{\prime}\subset D)=D^{\prime} and pr2(DD)=D.pr_{2}(D^{\prime}\subset D)=D. This defines a vector bundle α=(pr2)pr1(Lκ)\mathcal{I}^{\alpha}=(pr_{2})_{*}pr_{1}^{*}(L^{\kappa}). Denote the dual of α\mathcal{I}^{\alpha} by VαV^{\alpha}. Its locality structure will determine an embedding Πi=1nregα1𝑖(Vα)\Pi^{n}_{i=1}\mathbb{P}_{\mathcal{H}_{reg}^{\alpha}}^{1}\xrightarrow[]{i}\mathbb{P}(V^{\alpha}). Then one defines the space ZκαZ^{\alpha}_{\kappa} as the closure of the image of ii in (Vα)\mathbb{P}(V^{\alpha}).

As we briefly discussed above, when the matrix κ\kappa is the Cartan matrix of GG, Mirkovic showed that ZκαZ^{\alpha}_{\kappa} coincides with ZGαZ_{G}^{\alpha}888 This is done by showing (qα)Zκα(qα)ZGα=ZGα(q^{\alpha})^{*}Z^{\alpha}_{\kappa}\cong(q^{\alpha})^{*}Z_{G}^{\alpha}=Z_{G}^{\prime\alpha}. It seems natural to use the notation ZκαZ_{\kappa}^{\prime\alpha} for (qα)Zκα(q^{\alpha})^{*}Z^{\alpha}_{\kappa} as we do for ZGαZ_{G}^{\prime\alpha} but we will not use that.. In principle, the ”κ\kappa-approach” is more general than the Coulomb branch construction since we can consider any symmetric matrix that is not necessarily the (modified) incidence matrix of a quiver (possibly with loops). In the upcoming paper[Don], we will provide some examples where the projection Zκα𝜋C×IαZ^{\alpha}_{\kappa}\xrightarrow[]{\pi}\mathcal{H}^{\alpha}_{C\times I} is not flat when the matrix κ\kappa is not the incidence matrix of a quiver. Therefore, the merit of zastava for such κ\kappa is subject to skepticism.

The drawback of this approach is that ZκαZ_{\kappa}^{\alpha} is defined as a closure and we lack a direct description. Additionally, it is unclear how to define the generalized transverse slices of 𝒢(GKM)\mathcal{G}(G_{KM}). However, the advantage is that the growth structure [Mir23] is much easier to define.

1.5 Identification of two generalizations of Zastava

Now Let C=𝔸1C=\mathbb{A}^{1}. The main result of this paper is that the local (projective) space construction coincides with the Coulomb branch construction when κ\kappa is the incidence matrix κ(Q)\kappa(Q) of a quiver QQ. We will clarify the following parallel structures for Coulomb branch 𝐌Qα\mathbf{M}^{\alpha}_{Q} and local projective space Zκ(Q)αZ^{\alpha}_{\kappa(Q)}. In the table below, we omit the subscripts QQ and κ(Q)\kappa(Q). Also, we fix α=(ni)iI\alpha=(n_{i})_{i\in I}. Let G=iIGL(kni)G=\prod_{i\in I}GL(k^{n_{i}}) and TT be the diagonal subgroup of GG 999These G,TG,T depend on α\alpha and are different from the G,TG,T that appeared in the Zastava before. It is clear which G,TG,T we mean from the context.. Here, we omit α\alpha for the notations involving \mathcal{R}.

𝐌α\mathbf{M}^{\alpha} ZαZ^{\alpha}
projective embedding 𝐌αjαα(HG(1+))\mathbf{M}^{\alpha}\xhookrightarrow{j_{\alpha}}\mathbb{P}_{\mathcal{H}^{\alpha}}(H_{*}^{{G}}(\mathcal{R}^{+}_{\leq 1})^{*}) Zαiαα(Vα)Z^{\alpha}\xhookrightarrow{i_{\alpha}}\mathbb{P}_{\mathcal{H}^{\alpha}}(V^{\alpha})
regular fibers over reg\mathcal{H}_{reg} product of 1\mathbb{P}^{1}’s product of 1\mathbb{P}^{1}’s
global basis basis of HG(1+)H_{*}^{G}(\mathcal{R}^{+}_{\leq 1}) basis of Γ(α,α)\Gamma(\mathcal{H}^{\alpha},\mathcal{I}^{\alpha})

Here 1+\mathcal{R}^{+}_{\leq 1} is certain subspace of the positive part +\mathcal{R}^{+} of .\mathcal{R}. The embedding jαj_{\alpha} follows from Lemma 6, while the embedding iαi_{\alpha} is based on the definition of the local space. The space 𝐌α\mathbf{M}^{\alpha} is over Spec(HG(pt))α.Spec(H_{G}^{*}(pt))\cong\mathcal{H}^{\alpha}. Let Cα=iICni.C^{\alpha}=\prod_{i\in I}C^{n_{i}}.

Let us see how to identify 𝐌α\mathbf{M}^{\alpha} and ZαZ^{\alpha} after pulling back under CαqααC^{\alpha}\xrightarrow[]{q^{\alpha}}\mathcal{H}^{\alpha}.

Under the pull back qαq^{\alpha}, the space HG(1+)H_{*}^{G}(\mathcal{R}^{+}_{\leq 1}) becomes HT(1+)H_{*}^{T}(\mathcal{R}^{+}_{\leq 1}) and Γ(α,α)\Gamma(\mathcal{H}^{\alpha},\mathcal{I}^{\alpha}) becomes Γ(Cα,(qα)α).\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha}). We will identify HT(1+)relΓ(Cα,(qα)α)H_{*}^{T}(\mathcal{R}^{+}_{\leq 1})\xrightarrow[\cong]{rel}\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha}) in section 4.1, as free rank 1 modules101010In each component. over HT(1+)𝒪((qα)Gr(𝒯α)).H^{*}_{T}(\mathcal{R}^{+}_{\leq 1})\cong\mathcal{O}((q^{\alpha})^{*}Gr(\mathcal{T}^{\alpha})).

In component, the isomorphism HT(ϖβ)relΓ(Cα,(qα)α,β)H_{*}^{T}(\mathcal{R}_{\varpi_{\beta}})\xrightarrow[\cong]{rel}\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta}) is given by the ring isomorphism HT(ϖβ)𝒪((qα)Grβ(𝒯α))H_{*}^{T}(\mathcal{R}_{\varpi_{\beta}})\cong\mathcal{O}((q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha})) and choosing basis [ϖβ][\mathcal{R}_{\varpi_{\beta}}] of HT(ϖβ)H_{*}^{T}(\mathcal{R}_{\varpi_{\beta}}) and basis uβ1u^{\beta}\otimes 1 of Γ(Cα,(qα)α,β)\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta}) (see 4.1 for the notations).

By a simple lemma (Lemma 4.3), it suffices to prove that the dotted arrow below exists as an isomorphism, i.e., the isomorphism relrel restricts to an isomorphism of two embeddings over the regular part reg\mathcal{H}_{reg}(lemma 18).

(qα)Zregα{(q^{\alpha})^{*}Z_{reg}^{\alpha}}(qα)Zα{(q^{\alpha})^{*}Z^{\alpha}}Cα((qα)Vα){\mathbb{P}_{C^{\alpha}}((q^{\alpha})^{*}V^{\alpha})}(qα)𝐌regα{(q^{\alpha})^{*}\mathbf{M}_{reg}^{\alpha}}(qα)𝐌α{(q^{\alpha})^{*}\mathbf{M}^{\alpha}}Cα(HT(1+)){\mathbb{P}_{C^{\alpha}}(H_{*}^{T}(\mathcal{R}^{+}_{\leq 1})^{*})}iα\scriptstyle{i_{\alpha}}rel\scriptstyle{rel\simeq}jα\scriptstyle{j_{\alpha}}

To produce the dotted arrow, we will show their homogeneous coordinate rings are isomorphic. Any basis of HT(1+)H_{*}^{T}(\mathcal{R}^{+}_{\leq 1}) (resp. Γ(Cα,(qα)α)\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha})) over HT(1+)H^{*}_{T}(\mathcal{R}^{+}_{\leq 1}) (resp. 𝒪((qα)Gr(𝒯α))\mathcal{O}((q^{\alpha})^{*}Gr(\mathcal{T}^{\alpha})) generate the homogeneous coordinate rings of Cα(HT(1+))\mathbb{P}_{C^{\alpha}}(H_{*}^{T}(\mathcal{R}^{+}_{\leq 1})^{*}) (resp. Cα((qα)α)\mathbb{P}_{C^{\alpha}}((q^{\alpha})^{*}\mathcal{I}^{\alpha})) over 𝒪(Cα).\mathcal{O}(C^{\alpha}). There are canonical choices of basis for the localization over CregC_{reg} of both HT(1+)H_{*}^{T}(\mathcal{R}^{+}_{\leq 1}) and Γ(Cα,(qα)α)\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha}), which we will call the local bases111111There are no canonical choices of basis before localization. We call any of them a global basis. Here we use parallel terminologies for global and local basis but they have different natures. The local basis of HT(1+)H_{*}^{T}(\mathcal{R}^{+}_{\leq 1}) arises naturally from the localization theory of equivariant homology, and the local basis of Γ(Cα,(qα)α)\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha}) is inherent in the definition of local projective spaces. We then show that these are identified via relrel. Finally, we verify that they satisfy the same relations through direct computation, using our explicit understanding of multiplication formulas on both sides.

1.6 Background and further directions

The loop Grassmannians associated to quivers QQ contain a class of loop Grassmannians for affine groups, so it is a part of a current effort by mathematicians and physicists to lift features of Langlands program to dimension 2. This paper unifies approaches to 𝒢Q\mathcal{G}_{Q} via Coulomb branches and local spaces. For any quiver QQ the latter constructs Zastavas and Beilinson-Drinfeld Grassmannian [Mir23] (more generally, for symmetric integral matrices), and the former [BFN19] constructs Zastava and generalized slices.

This seems to make some features of Satake equivalence for general quiver QQ more accessible at the moment. This includes conjectural geometric constructions of the positive part U(𝔤Q+)U(\mathfrak{g}^{+}_{Q}) of the enveloping algebra for a Kac-Moody algebra associated with QQ (using Zastava) [FKM20, 5.4], and of its irreducible highest weight module V(λ)V(\lambda) (using generalized slices) in [BFN19, Conjecture 3.25]. Also, one hopes that the Geometric Casselman-Shalika theorem, i.e., the Whittaker version of Satake equivalence [Ras21, 6.36.1] can be extended to loop Grassmannian 𝒢Q\mathcal{G}_{Q} associated to quivers.

2 Mirković local projective space (Zastava)

We recall the notion of local space and local projective space introduced by Mirkovic[MYZ21, Mir23].

2.1 Hilbert scheme C×I\mathcal{H}_{C\times I}

Let II be a finite set. Let CC be a smooth algebraic curve over an algebraic closed field kk. We consider the Hilbert scheme C×I\mathcal{H}_{C\times I}, the moduli of finite subschemes of C×IC\times I. Let Ci=C×iC×IC_{i}=C\times i\subset C\times I and Cini\mathcal{H}^{n_{i}}_{C_{i}} be the Hilbert scheme of finite subschemes DD of length nin_{i} of CiC_{i}. For α=(ni)iI[I],\alpha=(n_{i})_{i\in I}\in\mathbb{N}[I], we define C×Iα=iICini.\mathcal{H}^{\alpha}_{C\times I}=\prod_{i\in I}\mathcal{H}^{n_{i}}_{C_{i}}. We understand α\alpha as a dimension vector.

We have C×I=α[I]C×Iα.\mathcal{H}_{C\times I}=\sqcup_{\alpha\in\mathbb{N}[I]}\mathcal{H}^{\alpha}_{C\times I}. We also call D=(Di)iIC×ID=(D_{i})_{i\in I}\in\mathcal{H}_{C\times I} an II-colored divisor where DiHCiD_{i}\in H_{C_{i}} is an effective divisor of CiC_{i} for iIi\in I. Since CC is a smooth algebraic curve, the Hilbert scheme Cn\mathcal{H}^{n}_{C} can also be viewed as the nn-th symmetric power SnCS^{n}C of CC, which is also the categorical quotient Cn//Sn.C^{n}//S_{n}.

2.2 The regular part of C×I\mathcal{H}_{C\times I}

For i,jIi,j\in I we define the (i,j)(i,j)-diagonal divisor ΔijC×I\Delta_{ij}\subset\mathcal{H}_{C\times I}. When iji\neq j the condition for DΔijD\in\Delta_{ij} is that the component divisors DiD_{i} and DjD_{j} meet. Also, Δii\Delta_{ii} is a discriminant divisor, a subscheme DC×ID\in\mathcal{H}_{C\times I} lies in Δii\Delta_{ii} if DiD_{i} is not discrete, i.e., some point has multiplicity >1>1. We call the complement of the union ΔijC×I\cup\Delta_{ij}\subset\mathcal{H}_{C\times I} the regular part of C×I\mathcal{H}_{C\times I} and denote it by reg\mathcal{H}_{reg}. We call a divisor DregD\in\mathcal{H}_{reg} a regular divisor. We say two divisors DD and DD^{\prime} are disjoint if they are disjoint after forgetting the colors.

2.3 Local line bundle LL on C×I\mathcal{H}_{C\times I}.

Definition 1.

A locality structure of a vector bundle 𝒱\mathcal{V} over C×I\mathcal{H}_{C\times I} is a system of isomorphisms : For two disjoint divisors121212Here, divisor means closed subschemes of C×I×SC\times I\times S that are flat over SS D,DC×ID,D^{\prime}\in\mathcal{H}_{C\times I},

𝒱D𝒱DiD,D𝒱DD\mathcal{V}_{D}\otimes\mathcal{V}_{D^{\prime}}\xrightarrow[]{i_{D,D^{\prime}}}\mathcal{V}_{D\cup D^{\prime}}

that satisfy the associative, commutative and unital properties.
Similarly, a local structure on a space YY over C×I\mathcal{H}_{C\times I} is a system of isomorphisms

YD×YDiD,DYDDY_{D}\times Y_{D^{\prime}}\xrightarrow[]{i_{D,D^{\prime}}}Y_{D\cup D^{\prime}}

that satisfy the same properties.

Remark 1.

The above structure (map between functors) is represented by the following. For any α=α1+α2\alpha=\alpha_{1}+\alpha_{2} , denote the vector bundle 𝒱\mathcal{V} on the component α\mathcal{H}^{\alpha} by 𝒱α.\mathcal{V}^{\alpha}.131313Here in the notation 𝒱α\mathcal{V}^{\alpha}, the superscript α\alpha indexes the connected component, not power. We have the addition of divisors

α1×α2aα1,α2α.\mathcal{H}^{\alpha_{1}}\times\mathcal{H}^{\alpha_{2}}\xrightarrow[]{a^{\alpha_{1},\alpha_{2}}}\mathcal{H}^{\alpha}.

Let pri,i=1,2pr_{i},i=1,2 and α1×α2priαi\mathcal{H}^{\alpha_{1}}\times\mathcal{H}^{\alpha_{2}}\xrightarrow[]{pr_{i}}\mathcal{H}^{\alpha_{i}} be the ii-th projection. Denote (α1×α2)disj(\mathcal{H}^{\alpha_{1}}\times\mathcal{H}^{\alpha_{2}})_{disj} by the open part of α1×α2\mathcal{H}^{\alpha_{1}}\times\mathcal{H}^{\alpha_{2}} that consists (D,D)(D,D^{\prime}) such that Dα1,Dα2D\in\mathcal{H}^{\alpha_{1}},D^{\prime}\in\mathcal{H}^{\alpha_{2}} are disjoint. A locality structure on a vector bundle 𝒱\mathcal{V} over C×I\mathcal{H}_{C\times I} is a system of isomorphisms over (α1×α2)disj(\mathcal{H}^{\alpha_{1}}\times\mathcal{H}^{\alpha_{2}})_{disj}

𝒱α1𝒱α2:=pr1𝒱α1pr2𝒱α2iα1,α2(aα1,α2)𝒱α,\mathcal{V}^{\alpha_{1}}\boxtimes\mathcal{V}^{\alpha_{2}}:=pr_{1}^{*}\mathcal{V}^{\alpha_{1}}\otimes pr_{2}^{*}\mathcal{V}^{\alpha_{2}}\xrightarrow[]{i^{\alpha_{1},\alpha_{2}}}(a^{\alpha_{1},\alpha_{2}})^{*}\mathcal{V}^{\alpha},

for any α1,α2\alpha_{1},\alpha_{2} that satisfy the associative, commutative and unital properties. Later, we will have vector bundles with the notation L,L,\mathcal{I} and VV, where LL means it is of rank 1, i.e. a line bundle.

2.4 Induced vector bundle \mathcal{I}

To a local line bundle LL over C×I\mathcal{H}_{C\times I}, we associate an induced vector bundle \mathcal{I} over C×I\mathcal{H}_{C\times I}.

Let 𝒯C×I×(C×I)\mathcal{T}\subset\mathcal{H}_{C\times I}\times(C\times I) be the tautological scheme over C×I\mathcal{H}_{C\times I}, where the fiber 𝒯D\mathcal{T}_{D} at DC×ID\in\mathcal{H}_{C\times I} is DC×I.D\subset C\times I. Let Gr(𝒯)Gr(\mathcal{T}) be the relative Hilbert scheme Gr(𝒯/C×I)Gr(\mathcal{T}/\mathcal{H}_{C\times I}) over C×I\mathcal{H}_{C\times I} such that the fiber at DC×ID\in\mathcal{H}_{C\times I} is Gr(D):=DGr(D):=\mathcal{H}_{D}, where D\mathcal{H}_{D} is the Hilbert scheme of all subschemes DDD^{\prime}\subset D.

Let 𝒯α\mathcal{T}^{\alpha} be the component of 𝒯\mathcal{T} over C×Iα\mathcal{H}^{\alpha}_{C\times I} and Grβ(𝒯α)Gr^{\beta}(\mathcal{T}^{\alpha}) be the component of Gr(𝒯/C×I)Gr(\mathcal{T}/\mathcal{H}_{C\times I}) whose fiber at DD is Grβ(D):=β(D),Gr^{\beta}(D):=\mathcal{H}^{\beta}(D), the Hilbert scheme of subschemes DDD^{\prime}\subset D of length β\beta.

The scheme Gr(𝒯)Gr(\mathcal{T}) is a self correspondence of C×I\mathcal{H}_{C\times I},

C×Ipr1Gr(𝒯)pr2C×I,\mathcal{H}_{C\times I}\xleftarrow{pr_{1}}Gr(\mathcal{T})\xrightarrow[]{pr_{2}}\mathcal{H}_{C\times I},

where pr1(DD)=Dpr_{1}(D^{\prime}\subset D)=D^{\prime} and pr2(DD)=D.pr_{2}(D^{\prime}\subset D)=D.

Lemma 1.

The map pr2pr_{2} is finite flat so \mathcal{I} is a vector bundle.

Proof.

For brevity, assume I={1}I=\{1\}. For p=(p1<<pk)p=(p_{1}<\cdots<p_{k}), let Grp(𝒯)Gr_{p}(\mathcal{T}) be the partial flag space of all filtrations D1DkDD_{1}\subset\cdots\subset D_{k}\subset D with |Di|=pi.|D_{i}|=p_{i}. The maps Grq(𝒯)Grp(𝒯)Gr_{q}(\mathcal{T})\xrightarrow[]{}Gr_{p}(\mathcal{T}) are clearly finite flat when qq is obtained from pp by adding ps+1p_{s+1} after some psp_{s}. Build a tower of maps from p={0}p=\{0\} to p=(1,,n)p=(1,\cdots,n). The claim follows by the property that if X𝑓YX\xrightarrow[]{f}Y is flat, Y𝑔ZY\xrightarrow[]{g}Z is flat if and only if xgfZx\xrightarrow[]{g\circ f}Z is flat. ∎

Now define the induced vector bundle =(pr2)pr1(L)\mathcal{I}=(pr_{2})_{*}pr_{1}^{*}(L) and its dual V=V=\mathcal{I}^{*}.

Lemma 2.

A local structure of LL canonically induces a local structure on VV

Proof.

The locality structure is naturally induced from its dual vector bundle so we check the locality structure for \mathcal{I}. Here we only check the case where D,DD,D^{\prime} are regular for later use. The general case is similar. We compute the fiber D\mathcal{I}_{D} for D=k=1npkD=\sqcup^{n}_{k=1}p_{k}

D=((pr2)(pr1)L)D=EDLE.\mathcal{I}_{D}=((pr_{2})_{*}(pr_{1})^{*}L)_{D}=\oplus_{E\subset D}L_{E}.

Now for disjoint divisors D,DD,D^{\prime},

DD(EDLE)(EDLE)ED,EDLELEiD,DEEDDLEEDD,\mathcal{I}_{D}\otimes\mathcal{I}_{D^{\prime}}\cong(\oplus_{E\subset D}L_{E})\otimes(\oplus_{E^{\prime}\subset D^{\prime}}L_{E^{\prime}})\cong\oplus_{E\subset D,E^{\prime}\subset D^{\prime}}L_{E}\otimes L_{E^{\prime}}\xrightarrow[]{i^{\mathcal{I}}_{D,D^{\prime}}}\oplus_{E\cup E^{\prime}\in D\cup D^{\prime}}L_{E\cup E^{\prime}}\cong\mathcal{I}_{D\cup D^{\prime}},

where iD,Di^{\mathcal{I}}_{D,D^{\prime}} is the direct sum of the locality isomorphisms of LL. We have the last isomorphism since D,DD,D^{\prime} are disjoint, any subscheme FDDF\subset D\cup D^{\prime} can be uniquely written as the union of EDE\subset D and EDE^{\prime}\subset D^{\prime}. ∎

2.5 Local projective space of a local vector bundle

To a local vector bundle VV over C×I\mathcal{H}_{C\times I} we will associated its local projective space loc(V)(V)\mathbb{P}^{loc}(V)\subset\mathbb{P}(V).

First, at a point D=pC×IC×ID=p\in C\times I\subset\mathcal{H}_{C\times I} the locality condition is empty, so we define loc(V)p=(Vp)\mathbb{P}^{loc}(V)_{p}=\mathbb{P}(V_{p}). Now over a regular divisor D=pkD=\sum p_{k}, as required by locality condition we define

loc(V)D=k=1n(Vpk).\mathbb{P}^{loc}(V)_{D}=\prod^{n}_{k=1}\mathbb{P}(V_{p_{k}}).

Then the locality structure of k=1nVpkVD\otimes^{n}_{k=1}V_{p_{k}}\cong V_{D} gives the Segre embedding

loc(V)D(V)D.\mathbb{P}^{loc}(V)_{D}\hookrightarrow\mathbb{P}(V)_{D}.

Now we have a subspace Pregloc(V){P}^{loc}_{reg}(V) in (V)\mathbb{P}(V) defined over reg.\mathcal{H}_{reg}. We define the local projective space loc(V)\mathbb{P}^{loc}(V) as the closure of regloc(V)\mathbb{P}^{loc}_{reg}(V) in (V).\mathbb{P}(V). By definition regloc(V)\mathbb{P}^{loc}_{reg}(V) is a local space over .\mathcal{H}.

Remark 2.

Now we represent the Segre embedding. Denote by ei[I]e_{i}\in\mathbb{N}[I] which is 11 in the ii-th coordinate and 0 in the others. Denote Cα=iICiniC^{\alpha}=\prod_{i\in I}C_{i}^{n_{i}} Let qαq^{\alpha} be the quotient map CαqααC^{\alpha}\xrightarrow[]{q^{\alpha}}\mathcal{H}^{\alpha}. For iI,1jnii\in I,1\leq j\leq n_{i}, let prijpr_{ij} be the compositions of Cαi-th projectionCinij-th projectionCiC^{\alpha}\xrightarrow[]{\text{i-th projection}}C^{n_{i}}_{i}\xrightarrow[]{\text{j-th projection}}C_{i}.

Let Cregα=(qα)1(regα)C^{\alpha}_{reg}=(q^{\alpha})^{-1}(\mathcal{H}^{\alpha}_{reg}) and Creg=α[I]CregαC_{reg}=\sqcup_{\alpha\in\mathbb{N}[I]}C^{\alpha}_{reg}. Recall we denote the restriction of V,V,\mathcal{I} to CiC_{i} by Vei,eiV^{e_{i}},\mathcal{I}^{e_{i}}. For α=(ni)iI\alpha=(n_{i})_{i\in I}, the local structure of \mathcal{I} defines the isomorphism over CregC_{reg}

iIj{1,ai}prijeiiα(qα)α.\bigotimes_{i\in I}\bigotimes_{j\in\{1,\cdots a_{i}\}}pr_{ij}^{*}\mathcal{I}^{e_{i}}\xleftarrow[]{i_{\alpha}}(q^{\alpha})^{*}\mathcal{I}^{\alpha}. (1)

and the Segre embedding over CregC_{reg} is the corresponding embedding

iIj{1,ai}prij(Vei)𝐢α(qα)(Vα).\prod_{i\in I}\prod_{j\in\{1,\cdots a_{i}\}}pr_{ij}^{*}\mathbb{P}(V^{e_{i}})\xhookrightarrow{\mathbf{i}^{*}_{\alpha}}(q^{\alpha})^{*}\mathbb{P}(V^{\alpha}). (2)

It is more convenient to work over CαC^{\alpha} other than α\mathcal{H}^{\alpha} when doing calculations. We will always pullback objects over α\mathcal{H}^{\alpha} by qαq^{\alpha}. In practice, we will consider the image of 𝐢α\mathbf{i}^{*}_{\alpha} and its closure over CαC^{\alpha} which descents to regloc(V)\mathbb{P}^{loc}_{reg}(V) and loc(V)\mathbb{P}^{loc}(V) respectively under qαq^{\alpha}.

Remark 3.

We compute the fiber VpV_{p} for pp a point. Since Gr(p)={,p}Gr(p)=\{\emptyset,p\}, we have Vp=LLpkLpV_{p}=L^{*}_{\emptyset}\oplus L^{*}_{p}\cong k\oplus L^{*}_{p} and (Vp)1.\mathbb{P}(V_{p})\cong\mathbb{P}^{1}. By locality, the fiber VDV_{D} at a regular divisor DD of length α\alpha is141414Since we just compute the fiber we can ignore the index II and only count the total length |α||\alpha| of DD.

VD=(DDLD)k2|α|.V_{D}=(\oplus_{D^{\prime}\subset D}L_{D^{\prime}})^{*}\cong k^{2^{|\alpha|}}.

Hence over regular divisors, the fiber of loc(Vα)\mathbb{P}^{loc}(V^{\alpha}) is a product of 1\mathbb{P}^{1}’s and the local space loc(Vα)\mathbb{P}^{loc}(V^{\alpha}) can be viewed as a degeneration of a product of 1\mathbb{P}^{1}’s in 2|α|1.\mathbb{P}^{2^{|\alpha|}-1}.

2.6 Zastava space ZκZ_{\kappa} of a symmetric matrix κ\kappa

We define a local line bundle over C×I\mathcal{H}_{C\times I} associated with a symmetric integral matrix κMI()\kappa\in M_{I}(\mathbb{Z}) as

Lκ=𝒪(κΔ),L_{\kappa}=\mathcal{O}_{\mathcal{H}}(-\kappa\Delta),

where κΔ\kappa\Delta is the divisor ijκijΔij\sum_{i\leq j}\kappa_{ij}\Delta_{ij} in C×I.\mathcal{H}_{C\times I}. On each connected component, we can embed LκL_{\kappa} into the sheaf of the fraction field of CC. The locality structure of LL is then the multiplication of rational functions.

For the local line bundle LκL_{\kappa}, denote the induced vector bundle by κ\mathcal{I}_{\kappa} and its dual by VκV_{\kappa}. We define Zκα=loc(Vκα).Z^{\alpha}_{\kappa}=\mathbb{P}^{loc}(V_{\kappa}^{\alpha}). Now we fix a matrix κ\kappa and drop κ\kappa from the notation. Denote the projection Zα𝜋α.Z^{\alpha}\xrightarrow[]{\pi}\mathcal{H}^{\alpha}. Abusing notation, we will also denote the pullback of π\pi under qαq^{\alpha} by π\pi.

2.7 ZαZ^{\alpha} for C=𝔸1C=\mathbb{A}^{1}

From now on, we set C=𝔸1C=\mathbb{A}^{1} to be the affine line. We fix a dimension vector α\alpha. Since the matrix κ\kappa is also fixed we write L=Lκ,=κ,V=Vκ.L=L_{\kappa},\mathcal{I}=\mathcal{I}_{\kappa},V=V_{\kappa}. Denote by 𝒮α(α)\mathcal{S}_{\mathcal{H}^{\alpha}}(\mathcal{I}^{\alpha}) the symmetric algebra generated by α\mathcal{I}^{\alpha} over α\mathcal{H}^{\alpha}. Let A0A_{0} be151515The notation A0A_{0} is only used in section 2.7. It does not reflect its dependence on α\alpha. Since we fix α\alpha in 2.7, this will not cause confusion. The same happens for notation 𝐈\mathbf{I}. its pullback under qα,q^{\alpha}, hence Proj(A0)(qα)(Vα).Proj(A_{0})\cong(q^{\alpha})^{*}\mathbb{P}(V^{\alpha}). Let A0,regA_{0,reg} be its localization over Cregα=(qα)regC^{\alpha}_{reg}=(q^{\alpha})^{*}\mathcal{H}_{reg}.
Let 𝐈reg\mathbf{I}_{reg} be an ideal in A0,regA_{0,reg} such that Proj(A0,reg/𝐈reg)(qα)Zα|reg.Proj(A_{0,reg}/\mathbf{I}_{reg})\cong(q^{\alpha})^{*}Z^{\alpha}|_{reg}.
Since α\mathcal{I}^{\alpha} is a free sheaf over α\mathcal{H}^{\alpha}, once we choose a trivialization of it, we can write the algebra A0A_{0} as a polynomial ring.

2.7.1 Explicit description of locality structure on \mathcal{I}

In this subsection we will study the ideal 𝐈reg\mathbf{I}_{reg} that defines the Segre embedding on Creg.C_{reg}.

Since we work over CαC^{\alpha} rather than α\mathcal{H}^{\alpha}, it is necessary to understand the pullback of the line bundle LL under qαq^{\alpha}. We fix a coordinate aa on C=A1C=A^{1} so get a coordinate system (aji)iI,j{1,,ni}(a^{i}_{j})_{i\in I,j\in\{1,\cdots,n_{i}\}} for CαC^{\alpha}.

Lemma 3.

The pullback of the divisors are

(qα)Δii=div(lj(aliaji), (qα)Δii=div(l,j(aliaji)).(q^{\alpha})^{*}\Delta_{ii}=div(\prod_{l\neq j}(a^{i}_{l}-a^{i}_{j}),\text{ }(q^{\alpha})^{*}\Delta_{ii^{\prime}}=div(\prod_{l,j}(a^{i}_{l}-a^{i^{\prime}}_{j})).
Proof.

The second formula is clear. This first follows from the case when I={1}I=\{1\} and α=2\alpha=2 (recall αI\alpha\in\mathbb{Z}^{I} is a dimension vector and when II is a one-point set, α)\alpha\in\mathbb{Z}). ∎

Denote the map replacing \mathcal{I} by LL in (1) by iαLi^{L}_{\alpha}. We will describe the map iαi_{\alpha} in (1) from the corresponding maps iαLi^{L}_{\alpha} of LL. Note in lemma 2, we described the locality structure of \mathcal{I} by morphisms between functor of points. Recall the map

pr1Gr(𝒯)pr2\mathcal{H}\xleftarrow[]{pr_{1}}Gr(\mathcal{T})\xrightarrow[]{pr_{2}}\mathcal{H}

On the component α,\mathcal{H}^{\alpha}, it is

0βαβpr1α0βαGrβ(𝒯α)pr2αα,\sqcup_{0\leq\beta\leq\alpha}\mathcal{H^{\beta}}\xleftarrow[]{pr^{\alpha}_{1}}\sqcup_{0\leq\beta\leq\alpha}Gr^{\beta}(\mathcal{T}^{\alpha})\xrightarrow[]{pr^{\alpha}_{2}}\mathcal{H}^{\alpha},

which is the disjoint union of

βpr1α,βGrβ(𝒯α)pr2α,βα,\mathcal{H^{\beta}}\xleftarrow[]{pr^{\alpha,\beta}_{1}}Gr^{\beta}(\mathcal{T}^{\alpha})\xrightarrow[]{pr^{\alpha,\beta}_{2}}\mathcal{H}^{\alpha}, (3)

Let α,β=(pr2α,β)(pr1α,β)L.\mathcal{I}^{\alpha,\beta}=(pr^{\alpha,\beta}_{2})_{*}(pr^{\alpha,\beta}_{1})^{*}L. For α\alpha we define an II-colored index set Sα=iI{1,,ni}.S^{\alpha}=\sqcup_{i\in I}\{1,\cdots,n_{i}\}. For any subset SSα,S\subset S^{\alpha}, let SiS_{i} be the ii-th component of SS and let |S|=(|Si|)iI.|S|=(|S_{i}|)_{i\in I}. Define CS=iIjSiCijC^{S}=\prod_{i\in I}\prod_{j\in S_{i}}C_{ij}, where Cij=CC_{ij}=C so that Cα=CSα.C^{\alpha}=C^{S^{\alpha}}. Now we will see what (qα)α,β(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta} is. We introduce the following

βpr3α,β(qα)Grβ(𝒯α)pr4α,β:=(qα)(pr2α,β)Cα,\mathcal{H}^{\beta}\xleftarrow[]{pr_{3}^{\alpha,\beta}}(q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha})\xrightarrow[]{pr^{\alpha,\beta}_{4}:=(q^{\alpha})^{*}(pr_{2}^{\alpha,\beta})}C^{\alpha}, (4)

where pr3α,βpr_{3}^{\alpha,\beta} is the composition of maps from (qα)Grβ(𝒯α)(q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha}) to Grβ(𝒯α)Gr^{\beta}(\mathcal{T}^{\alpha}) and Grβ(𝒯α)pr1α,ββGr^{\beta}(\mathcal{T}^{\alpha})\xrightarrow[]{pr^{\alpha,\beta}_{1}}\mathcal{H}^{\beta}. We have

(qα)α,β(pr4α,β)(pr3α,β)(L).(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta}\cong(pr^{\alpha,\beta}_{4})_{*}(pr^{\alpha,\beta}_{3})^{*}(L). (5)

Over Cregα,C^{\alpha}_{reg}, (qα)Grβ(𝒯regα)S,|S|=|β|Cregα(q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha}_{reg})\cong\sqcup_{S,|S|=|\beta|}C^{\alpha}_{reg} and the map pr1α,βpr_{1}^{\alpha,\beta} factors through S,|S|=|β|Cregβ\sqcup_{S,|S|=|\beta|}C^{\beta}_{reg} so (4) becomes

regβqβS,|S|=|β|Cregβprα,SS,|S|=|β|CregαidCregα,\mathcal{H}_{reg}^{\beta}\xleftarrow[]{\sqcup q^{\beta}}\sqcup_{S,|S|=|\beta|}C^{\beta}_{reg}\xleftarrow[]{\sqcup pr^{\alpha,S}}\sqcup_{S,|S|=|\beta|}C^{\alpha}_{reg}\xrightarrow[]{\sqcup id}C^{\alpha}_{reg}, (6)

where prα,Spr^{\alpha,S} is the projection of the SS-components of CαC^{\alpha} to CSC^{S} composed with the identification map CSCβC^{S}\cong C^{\beta} (and we still denote by the same notation prα,Spr^{\alpha,S} when restricting on the regular part). On the SS-component, this is

regβqβCregβprα,SCregα=Cregα.\mathcal{H}_{reg}^{\beta}\xleftarrow[]{q^{\beta}}C^{\beta}_{reg}\xleftarrow[]{pr^{\alpha,S}}C^{\alpha}_{reg}=C^{\alpha}_{reg}. (7)

Denote the pullback of LregβL_{reg}^{\beta} under prα,Sqβpr^{\alpha,S}\circ q^{\beta} by Lregα,SL_{reg}^{\alpha,S}. Hence (qα)α|reg(q^{\alpha})^{*}\mathcal{I}^{\alpha}|_{reg} is the direct sum of line bundles Lregα,SL_{reg}^{\alpha,S}, i.e.

(qα)regα0βα(qα)regα,β0βαS,|S|=|β|Lregα,S=S,SSαLregα,S.(q^{\alpha})^{*}\mathcal{I}^{\alpha}_{reg}\cong\oplus_{0\leq\beta\leq\alpha}(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta}_{reg}\cong\oplus_{0\leq\beta\leq\alpha}\oplus_{S,|S|=|\beta|}L^{\alpha,S}_{reg}=\oplus_{S,S\subset S^{\alpha}}L^{\alpha,S}_{reg}. (8)

Here when β=0\beta=0, S=S=\emptyset and Lα,L^{\alpha,\emptyset} is the pullback from C0C^{0} to CαC^{\alpha}, which is the trivial line bundle (structure sheaf). Apply (8) to the case where α=ei\alpha=e_{i}, since qei=idq^{e_{i}}=id and Cregei=CeiC^{e_{i}}_{reg}=C^{e_{i}}, we have

eiLei,Lei.\mathcal{I}^{e_{i}}\cong L^{e_{i},\emptyset}\oplus L^{e_{i}}.

Pulling back under CαprijCeiC^{\alpha}\xrightarrow[]{pr_{ij}}C^{e_{i}}, we have

prijeiLα,prijLei.pr_{ij}^{*}\mathcal{I}^{e_{i}}\cong L^{\alpha,\emptyset}\oplus pr_{ij}^{*}L^{e_{i}}.

Now the locality isomorphism iαi_{\alpha} of \mathcal{I} in (1) (to be described) becomes

(iIj(Sα)i(Lα,prijLei))|CregαiαS,SSαLregα,S.(\otimes_{i\in I}\otimes_{j\in(S_{\alpha})_{i}}(L^{\alpha,\emptyset}\oplus pr_{ij}^{*}L^{e_{i}}))|_{C^{\alpha}_{reg}}\xleftarrow[]{i_{\alpha}}\oplus_{S,S\subset S_{\alpha}}L^{\alpha,S}_{reg}.

Expanding the tensor product, since Lα,L^{\alpha,\emptyset}\otimes\mathcal{L} is the trivial line bundle, for any SSαS\subset S^{\alpha} the SS-component of the map iαi_{\alpha} is denoted by iαSi^{S}_{\alpha}

iIjSiprijLei|CregαiαSLregα,S.\otimes_{i\in I}\otimes_{j\in S_{i}}pr_{ij}^{*}L^{e_{i}}|_{C^{\alpha}_{reg}}\xleftarrow[]{i_{\alpha}^{S}}L_{reg}^{\alpha,S}. (9)

Finally, we can describe the isomorphism iαSi^{S}_{\alpha}. It is the pullback under CregβprSCregαC^{\beta}_{reg}\xleftarrow[]{pr^{S}}C^{\alpha}_{reg} for β=|S|\beta=|S| of an open part161616In remark (1), we decompose β\beta completely. of the locality structure of LβL^{\beta}

iIj(Sβ)iprij(Lei)|CregβiβL(qβ)Lβ|Cregβ.\otimes_{i\in I}\otimes_{j\in(S_{\beta})_{i}}pr^{*}_{ij}(L^{e_{i}})|_{C^{\beta}_{reg}}\xleftarrow[\cong]{i^{L}_{\beta}}(q^{\beta})^{*}L^{\beta}|_{C^{\beta}_{reg}}. (10)

It is not hard to see that iαi_{\alpha} as described above are the isomorphisms that represent the locality isomorphisms in lemma 2.

2.7.2 Bases of global section Γ(Cα,(qα))\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}) as 𝒪(Cα)\mathcal{O}(C^{\alpha})-module (global bases)

For each β\beta, we fix a basis sβΓ(Cβ,(qβ)Lβ)s^{\beta}\in\Gamma(C^{\beta},(q^{\beta})^{*}L^{\beta}) adapted to the above local structure as follows. First we fix a basis sis_{i} of global section of LeiL^{e_{i}} over ei.\mathcal{H}^{e_{i}}. For any β=(ki)iI[I],\beta=(k_{i})_{i\in I}\in\mathbb{N}[I], let

sβ=l(β)iβ(iIj(Sβ)iprijsi)Γ(Cregβ,(qβ)Lβ|Creg),s^{\beta}=l(\beta)i_{\beta}(\otimes_{i\in I}\otimes_{j\in(S_{\beta})_{i}}pr^{*}_{ij}s_{i})\in\Gamma(C^{\beta}_{reg},(q^{\beta})^{*}L^{\beta}|_{C_{reg}}), (11)

where 171717Here, we choose a basis according to κ\kappa, later, we will define lQ(β)l_{Q}(\beta) depending on the quiver QQ.

l(β)=i<i,l(Sβ)i,j(Sβ)i(aliaji)κiiiI,lj,l,j(Sβ)i(aliaji)κiil(\beta)=\prod_{i<i^{\prime},l\in(S_{\beta})_{i},j\in(S_{\beta})_{i^{\prime}}}(a^{i}_{l}-a^{i^{\prime}}_{j})^{\kappa_{ii^{\prime}}}\prod_{i\in I,l\neq j,l,j\in(S_{\beta})_{i}}(a^{i}_{l}-a^{i}_{j})^{\kappa_{ii}} (12)

is in the fraction field 𝒦(β)\mathcal{K}(\mathcal{H}^{\beta}) of β.\mathcal{H}^{\beta}. Here we still denote by iβi_{\beta} the induced map between global sections of line bundles in (10). By lemma 3, the section sβs^{\beta} extend across the diagonal to a section of Cβ.C^{\beta}. Let WβW^{\beta} be the group such that Cβ/WββC^{\beta}/W^{\beta}\cong\mathcal{H}^{\beta}. It is easy to check that sβs^{\beta} is WβW^{\beta}-invariant. Let uβΓ(β,Lβ)u^{\beta}\in\Gamma(\mathcal{H}^{\beta},L^{\beta}) be the descent of sβs^{\beta} under qβ.q^{\beta}.

Let MβM^{\beta} be the global sections of LβL^{\beta} over β\mathcal{H}^{\beta}. It is a rank 1 free 𝒪(β)\mathcal{O}(\mathcal{H}^{\beta})-module generated by uβu^{\beta}. By (5),

Γ(Cα,(qα)α,β)=Mβ𝒪(β)𝒪((qα)Grβ(𝒯α))\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta})=M^{\beta}\otimes_{\mathcal{O}(\mathcal{H}^{\beta})}\mathcal{O}((q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha}))

as O(Cα)O(C^{\alpha})-module. By choosing any basis {ypα,β}\{y^{\alpha,\beta}_{p}\} of 𝒪((qα)Grβ(𝒯α))\mathcal{O}((q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha})) as 𝒪(Cα)\mathcal{O}(C^{\alpha})-module, which is chosen as a basis of 𝒪(Grβ(𝒯α))\mathcal{O}(Gr^{\beta}(\mathcal{T}^{\alpha})) as 𝒪(α)\mathcal{O}(\mathcal{H}^{\alpha})-module, we have the basis {zpα,β=uβyp}\{z^{\alpha,\beta}_{p}=u^{\beta}\otimes y_{p}\} of Γ(Cα,(qα)α,β)\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta}). When α\alpha is fixed, we abbreviate zpα,βz^{\alpha,\beta}_{p} as zpβz^{\beta}_{p}. Such a basis is called a global basis.

2.7.3 The basis of Γ(Cregα,(qα)reg)\Gamma(C^{\alpha}_{reg},(q^{\alpha})^{*}\mathcal{I}_{reg}) (local basis)

From 1𝒪((qα)Grβ(α))1\in\mathcal{O}((q^{\alpha})^{*}Gr^{\beta}(\mathcal{H}^{\alpha})), we get an element uβ1u^{\beta}\otimes 1 in

Γ(Cα,(qα)α,β)=Mβ𝒪(β)𝒪((qα)Grβ(𝒯α)).\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta})=M^{\beta}\otimes_{\mathcal{O}(\mathcal{H}^{\beta})}\mathcal{O}((q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha})).

Restricting uβ1u^{\beta}\otimes 1 to CregαC^{\alpha}_{reg}, by (6), we have sβ|reg1s^{\beta}|_{reg}\otimes 1 in

Γ(Cregα,(qα)regα,β)=S,|S|=βΓ(Cregα,Lregα,S)=S,|S|=βΓ(Cregβ,(qβ)Lregβ)𝒪(Cregβ)𝒪(Cregα).\Gamma(C_{reg}^{\alpha},(q^{\alpha})^{*}\mathcal{I}_{reg}^{\alpha,\beta})=\bigoplus_{S,|S|=\beta}\Gamma(C_{reg}^{\alpha},L^{\alpha,S}_{reg})=\bigoplus_{S,|S|=\beta}\Gamma(C_{reg}^{\beta},(q^{\beta})^{*}L^{\beta}_{reg})\otimes_{\mathcal{O}(C^{\beta}_{reg})}{\mathcal{O}(C^{\alpha}_{reg})}.

In the decomposition in (8), denote the SS-component of sβ|reg1s^{\beta}|_{reg}\otimes 1 by sα,SΓ(Cregα,Lregα,S)s^{\alpha,S}\in\Gamma(C^{\alpha}_{reg},L^{\alpha,S}_{reg}). We will refer to sα,Ss^{\alpha,S} as the local basis. Now we fix an α\alpha and abbreviate sα,Ss^{\alpha,S} as sS.s^{S}. By our choices of bases sβs^{\beta} in (11) and the relation between (9) and (10), we have

sS=l(S)iαS(iIjSiprijsi),s^{S}=l(S)i^{S}_{\alpha}(\otimes_{i\in I}\otimes_{j\in S_{i}}pr^{*}_{ij}s_{i}), (13)

where

l(S)=i<i,lSi,jSi(aliaji)κiiiI,lj,l,jSi(aliaji)κiil(S)=\prod_{i<i^{\prime},l\in S_{i},j\in S_{i^{\prime}}}(a^{i}_{l}-a^{i^{\prime}}_{j})^{\kappa_{ii^{\prime}}}\prod_{i\in I,l\neq j,l,j\in S_{i}}(a^{i}_{l}-a^{i}_{j})^{\kappa_{ii}} (14)

is the pullback of l(β)l(\beta) under Cregβprα,SCregαC^{\beta}_{reg}\xleftarrow[]{pr^{\alpha,S}}C^{\alpha}_{reg} (see (7)) and we still denote by iαSi^{S}_{\alpha} the induced map between global sections of line bundles in (9). By definition {sS,SSα}\{s^{S},S\subset S^{\alpha}\} generate the algebra A0,regA_{0,reg} over 𝒪(Cregα)\mathcal{O}(C_{reg}^{\alpha}). Now we can describe the ideal 𝐈reg\mathbf{I}_{reg} in terms of local basis sSs^{S}.

Lemma 4.

The ideal 𝐈reg\mathbf{I}_{reg} is generated by

sAl(A)sBl(B)=sABl(AB)sABl(AB)\frac{s^{A}}{l(A)}\frac{s^{B}}{l(B)}=\frac{s^{A\cup B}}{l(A\cup B)}\frac{s^{A\cap B}}{l(A\cap B)}

for all A,BSα.A,B\subset S_{\alpha}.

Proof.

This is a standard result of Segre embedding. Since we cannot find a reference, we give a proof in the appendix. ∎

2.7.4 Transition between a global basis zpβz^{\beta}_{p} and the local basis sSs^{S} over the regular part

First, we state a simple lemma about the module structure under localization.

Lemma 5.

The following diagram commutes
𝒪(Grβ(𝒯α))×Γ(Cα,(qα)α,β){\mathcal{O}(Gr^{\beta}(\mathcal{T}^{\alpha}))\times\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta})}Γ(Cα,(qα)α,β){\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta})}𝒪(Grβ(𝒯regα))×Γ(Cregα,(qα)regα,β){\mathcal{O}(Gr^{\beta}(\mathcal{T}_{reg}^{\alpha}))\times\Gamma(C_{reg}^{\alpha},(q^{\alpha})^{*}\mathcal{I}_{reg}^{\alpha,\beta})}Γ(Cregα,(qα)regα,β){\Gamma(C_{reg}^{\alpha},(q^{\alpha})^{*}\mathcal{I}_{reg}^{\alpha,\beta})}S,|S|=|β|𝒪(Cregα)×S,|S|=βΓ(Cregα,Lregα,S){\bigoplus_{S,|S|=|\beta|}\mathcal{O}(C^{\alpha}_{reg})\times\bigoplus_{S,|S|=\beta}\Gamma(C_{reg}^{\alpha},L^{\alpha,S}_{reg})}S,|S|=βΓ(Cregα,Lregα,S),{\bigoplus_{S,|S|=\beta}\Gamma(C_{reg}^{\alpha},L^{\alpha,S}_{reg}),}loc×loc\scriptstyle{loc\times loc}loc\scriptstyle{loc}\scriptstyle{\cong}\scriptstyle{\cong}
where the horizontal maps are the ring actions on the modules.

Proof.

Clear from the definitions. ∎

Now we will study the following map in more details.

𝒪(Grβ(𝒯α))|reg𝒪(Grβ(𝒯regα))S,SSα,|S|=|β|𝒪(regα).\mathcal{O}(Gr^{\beta}(\mathcal{T}^{\alpha}))\xrightarrow[]{|_{reg}}\mathcal{O}(Gr^{\beta}(\mathcal{T}_{reg}^{\alpha}))\cong\bigoplus_{S,S\subset S_{\alpha},|S|=|\beta|}\mathcal{O}(\mathcal{H}^{\alpha}_{reg}). (15)

Let β=iIkii\beta=\sum_{i\in I}k_{i}i. Denote fs()f^{s}() the ss-th elementary symmetric functions on the variables inside the parenthesis.

Theorem 1.

a)

𝒪(Grβ(𝒯α))iI𝒪(Grki(𝒯ni)).\mathcal{O}(Gr^{\beta}(\mathcal{T}^{\alpha}))\cong\bigotimes_{i\in I}\mathcal{O}(Gr^{k_{i}}(\mathcal{T}^{n_{i}})).

b) The map (pr1,m)(pr_{1},m)

Grki(𝒯ni)(pr1,m)ki×niki,Gr^{k_{i}}(\mathcal{T}^{n_{i}})\xrightarrow[]{(pr_{1},m)}\mathcal{H}^{k_{i}}\times\mathcal{H}^{n_{i}-k_{i}},

where m((D,D)=DDm((D^{\prime},D)=D-D^{\prime} is an isomorphism.
c) The projection Grki(𝒯ni)pr2niGr^{k_{i}}(\mathcal{T}^{n_{i}})\xrightarrow[]{pr_{2}}\mathcal{H}^{n_{i}} is finite flat and 𝒪(Grki(𝒯ni))\mathcal{O}(Gr^{k_{i}}(\mathcal{T}^{n_{i}})) is an algebra extension over 𝒪(ni)\mathcal{O}(\mathcal{H}^{n_{i}}) generated by formal variables cli,dji,1lki,1jnikic^{i}_{l},d^{i}_{j},1\leq l\leq k_{i},1\leq j\leq n_{i}-k_{i} subject to the relations

<1lki,l+j=siclidjifsi(a1i,,anii)><\sum_{1\leq l\leq k_{i},l+j=s_{i}}c^{i}_{l}d^{i}_{j}-f^{s_{i}}(a^{i}_{1},\cdots,a^{i}_{n_{i}})>

for all sis_{i} such that 1sini.{1\leq s_{i}\leq n_{i}}.
d) Under the above isomorphism and (15), the image of clic^{i}_{l} in the SiS_{i} component is fl(ari)rSif^{l}(a^{i}_{r})_{r\in S_{i}} and the image of djid^{i}_{j} in the SiS_{i} component is fj(ari)r{1,,ni}Sif^{j}(a^{i}_{r})_{r\in\{1,\cdots,n_{i}\}\setminus S_{i}}.

Proof.

a) Since

(Grβ(𝒯α))iI(Grki(𝒯ni)).(Gr^{\beta}(\mathcal{T}^{\alpha}))\cong\prod_{i\in I}(Gr^{k_{i}}(\mathcal{T}^{n_{i}})).

b) Clear.
c) We omit the index ii. The map pr2pr_{2} being finite flat is already proved. Since we have the map Grk(𝒯)pr1kGr^{k}(\mathcal{T})\xrightarrow[]{pr_{1}}\mathcal{H}^{k}, where pr1(D,D)=Dpr_{1}(D^{\prime},D)=D, we can pull back functions on k\mathcal{H}^{k} to Grk(𝒯)Gr^{k}(\mathcal{T}). Let 𝒮h\mathcal{S}_{h} be the symmetric group of order hh. We have fl(a1,,ak)k[a1,,ak]𝒮k𝒪(k).f^{l}(a_{1},\cdots,a_{k})\in k[a_{1},\cdots,a_{k}]^{\mathcal{S}_{k}}\cong\mathcal{O}(\mathcal{H}^{k}). Let their pullback under pr1pr_{1} be c1,ckc_{1},\cdots c_{k}. We also have a map Grk(𝒯)𝑚nkGr^{k}(\mathcal{T})\xrightarrow[]{m}\mathcal{H}^{n-k} where m(D,D)=DDm(D^{\prime},D)=D-D^{\prime}. Similarly, we have fj(a1,,ank)k[a1,,ank]𝒮nk𝒪(nk).f^{j}(a_{1},\cdots,a_{n-k})\in k[a_{1},\cdots,a_{n-k}]^{\mathcal{S}_{n-k}}\cong\mathcal{O}(\mathcal{H}^{n-k}). and let their pullback under mm be d1,dnkd_{1},\cdots d_{n-k}. For (D,D)Gr(𝒯)(D^{\prime},D)\in Gr(\mathcal{T}), where D=t=1npt,D=tSptD=\sum^{n}_{t=1}p_{t},D^{\prime}=\sum_{t\in S}p_{t}, We have

l+j=scldj((D,D))=l+j=scl(D)dj(DD)\sum_{l+j=s}c_{l}d_{j}((D^{\prime},D))=\sum_{l+j=s}c_{l}(D)d_{j}(D-D^{\prime})\\
=l+j=sfl((ar)rS)fj((at)t{1,,a}S)=fs(a1,,an).=\sum_{l+j=s}f^{l}((a_{r})_{r\in S})f^{j}((a_{t})_{t\in\{1,\cdots,a\}\setminus S})=f^{s}(a_{1},\cdots,a_{n}).

Since fs(a1,,an)f^{s}(a_{1},\cdots,a_{n}) is 𝒮n\mathcal{S}^{n} invariant, it can be viewed as functions on n\mathcal{H}_{n} of the same form pulling back by pr2pr_{2}.
By part(b), the algebra 𝒪(Grk(𝒯n))\mathcal{O}(Gr^{k}(\mathcal{T}^{n})) is already generated by cl,djc_{l},d_{j}. Over reg\mathcal{H}_{reg} since pr2pr_{2} is finite flat, the degree can be seen over reg\mathcal{H}_{reg} and from (15) it is (nk).\tbinom{n}{k}. Now, these relations exhaust all from the well-known algebraic fact that the degree of the algebra extension is (nk).\tbinom{n}{k}.
d) Clear from the definition. ∎

Restricting to the regular part (as 𝒪(Creg)\mathcal{O}(C_{reg})-module), we still denote by {zpβ}\{z^{\beta}_{p}\} as a basis of the 𝒪(Creg)\mathcal{O}(C_{reg})-module Γ(Cregα,(qα)regα,β).\Gamma(C_{reg}^{\alpha},(q^{\alpha})^{*}\mathcal{I}_{reg}^{\alpha,\beta}).

3 Coulomb branch

We briefly review the mathematical definition of the Coulomb branch[BFN18].

Let GG be a reductive group and NN be a representation of GG. Let 𝒦=((t))\mathcal{K}=\mathbb{C}((t)) and 𝒪=[[t]]\mathcal{O}=\mathbb{C}[[t]]. We denote by 𝒢(G)=G𝒦/G𝒪\mathcal{G}(G)=G_{\mathcal{K}}/G_{\mathcal{O}} the loop Grassmannian of GG. Since GG acts on NN, G𝒪G_{\mathcal{O}} acts on N𝒪N_{\mathcal{O}} and we get an associated bundle G𝒦×G𝒪N𝒪G_{\mathcal{K}}\times_{G_{\mathcal{O}}}N_{\mathcal{O}}. It has an embedding ii into 𝒢(G)×N𝒦\mathcal{G}(G)\times N_{\mathcal{K}} given by i((g,v)¯)=(g¯,gv)i(\overline{(g,v)})=(\overline{g},g\cdot v). Let (G,N)\mathcal{R}_{(G,N)} be the preimage of 𝒢(G)×N𝒪\mathcal{G}(G)\times N_{\mathcal{O}} under ii. In this paper, we will call Borel-Moore homology just homology. In [BFN18], they define an algebra structure on the G𝒪G_{\mathcal{O}}-equivariant homology of \mathcal{R} and define the Coulomb branch of the pair (G,N)(G,N) to be the Spec of this algebra.

3.1 Localization

Let TGT\subset G be a Cartan subgroup and 𝔱\mathfrak{t} be its Lie algebra. Let T,NT\mathcal{R}_{T,N_{T}} be the variety of triples for the pair (T,NT)(T,N_{T}), where NTN_{T} is the NN considered as a representation of TT. Denote the localization of 𝔱\mathfrak{t} at all roots of GG and weights of NTN_{T} by 𝔱reg.\mathfrak{t}_{reg}. Let T,NT𝜄\mathcal{R}_{T,N_{T}}\xrightarrow[]{\iota}\mathcal{R} be the embedding, the pushforward homomorphism ι\iota_{*}

HT𝒪(T,NT)ιHT𝒪()H^{T_{\mathcal{O}}}_{*}(\mathcal{R}_{T,N_{T}})\xrightarrow[]{\iota_{*}}H^{T_{\mathcal{O}}}_{*}(\mathcal{R}) (16)

gives an algebra homomorphism, which becomes an isomorphism over 𝔱reg\mathfrak{t}_{reg}[BFN18, Lemma (5.17)].181818The notation 𝔱reg\mathfrak{t}_{reg} ignores its dependence on NN.

3.2 Formula for HT𝒪(T)H^{T_{\mathcal{O}}}_{*}(\mathcal{R}_{T})

Now we fix a representation NN and abbreviate NT,T\mathcal{R}_{N_{T},T} by T\mathcal{R}_{T}. We have T𝑝𝒢(T)\mathcal{R}_{T}\xrightarrow[]{p}\mathcal{G}(T) . Since we consider the homology, we only count the reduced part of 𝒢(T)\mathcal{G}(T) so we denote the reduced part of 𝒢(T)\mathcal{G}(T) (still) by 𝒢(T)\mathcal{G}(T). Then 𝒢(T)\mathcal{G}(T) is disjoint union of points tχt^{\chi}, where χ\chi are cocharacters of TT. The fiber of pp over tχt^{\chi} is a vector space, which we denote by VχV_{\chi}. So we have T=χVχ.\mathcal{R}_{T}=\sqcup_{\chi}V_{\chi}. Denote the T𝒪T_{\mathcal{O}}-equivariant fundamental class of VχV_{\chi} by rχr^{\chi}. Let ξi\xi_{i} be the characters of TT that appear in the representation NN. Denote by ξi(χ)\xi_{i}(\chi) the pairing of ξi\xi_{i} and χ\chi. For two integers k,lk,l, let us set

d(k,l)={0if k and l have the same signs.min(|k|,|l|)if k and l have different signs.d(k,l)=\left\{\begin{aligned} &0&\text{if }k\text{ and }l\text{ have the same signs.}\\ &min(|k|,|l|)&\text{if }k\text{ and }l\text{ have different signs.}\\ \end{aligned}\right.
Theorem 2.

[BFN18, Theorem 4.1] The algebra HT𝒪(T)H^{T_{\mathcal{O}}}_{*}(\mathcal{R}_{T}) is generated by rχr^{\chi} for all χX(T)\chi\in X_{*}(T) over HT𝒪(pt)H^{T_{\mathcal{O}}}_{*}(pt).
For two cocharacters λ\lambda and μ\mu, the multiplication of rλr^{\lambda} and rμr^{\mu} is

rλrμ=i=1nξid(ξi(λ),ξi(μ))rλ+μ.r^{\lambda}r^{\mu}=\prod^{n}_{i=1}\xi_{i}^{d(\xi_{i}(\lambda),\xi_{i}(\mu))}r^{\lambda+\mu}. (17)
Remark 4.

The coefficient before rλ+μr^{\lambda+\mu} depends on λ,μ\lambda,\mu and the representation NN. Denote the Grothendieck group of TT by K0(Rep(T))K^{0}(Rep(T)) and the monomials of X(T)X_{*}(T) by Mon(X(T))Mon(X_{*}(T)). For fixed λ,μ\lambda,\mu, the map K0(Rep(T))coλ,μMon(X(T))K^{0}(Rep(T))\xrightarrow[]{co^{\lambda,\mu}}Mon(X_{*}(T)) given by coλ,μ(N)=i=1nξid(ξi(λ),ξi(μ))co^{\lambda,\mu}(N)=\prod^{n}_{i=1}\xi_{i}^{d(\xi_{i}(\lambda),\xi_{i}(\mu))} is a homomorphism of monoids, i.e. for N=N1N2N=N_{1}\oplus N_{2} , we have

coλ,μ(N)=coλ,μ(N1)coλ,μ(N2).co^{\lambda,\mu}(N)=co^{\lambda,\mu}(N_{1})co^{\lambda,\mu}(N_{2}).

3.3 Compactified Coulomb branch 𝐌Qα\mathbf{M}^{\alpha}_{Q} of quiver gauge theory

Here we consider a special case where the pair (G,N)(G,N) is given from a quiver QQ. Let Q=(I,E)Q=(I,E) be the quiver where II is the set of vertices and EE is the set of arrows. Given α=iInii[I]\alpha=\sum_{i\in I}n_{i}i\in\mathbb{N}[I], we view it as a dimension vector α=(ni)iI\alpha=(n_{i})_{i\in I}. Let Vi=kniV_{i}=k^{n_{i}} and V=(Vi)iIV=(V_{i})_{i\in I} is an II- graded vector space. Let N=ijHom(Vi,Vj)N=\oplus_{i\xrightarrow[]{}j}Hom(V_{i},V_{j}) and G=iGL(Vi)G=\prod_{i}GL(V_{i}) act on NN by conjugation.

Following [BFN18] 3(ii), for any vector space UU, we define 𝒢+(GL(U))𝒢((GL(U))\mathcal{G}^{+}(GL(U))\subset\mathcal{G}((GL(U)) as the moduli space of vector bundles 𝒰\mathcal{U} on the formal disc DD with a trivialization σ:𝒰DU𝒪D\sigma:\mathcal{U}_{D^{*}}\cong U\otimes\mathcal{O}_{D^{*}} on the punctured disc that extends through the puncture as an embedding σ:𝒰DU𝒪D\sigma:\mathcal{U}_{D}\hookrightarrow U\otimes\mathcal{O}_{D^{*}}. Let 𝒢GL(V)+=iI𝒢GL(Vi)+\mathcal{G}^{+}_{GL(V)}=\prod_{i\in I}\mathcal{G}^{+}_{GL(V_{i})}. Define +\mathcal{R}^{+} as the preimage of 𝒢GL(V)+𝒢GL(V)\mathcal{G}^{+}_{GL(V)}\subset\mathcal{G}_{GL(V)} under 𝒢GL(V)\mathcal{R}\xrightarrow[]{}\mathcal{G}_{GL(V)}. The homology group HGL(V)𝒪(+)H^{GL(V)_{\mathcal{O}}}_{*}(\mathcal{R}^{+}) forms a convolution subalgebra of HGL(V)𝒪()H^{GL(V)_{\mathcal{O}}}_{*}(\mathcal{R}).

In general, given a 0\mathbb{Z}^{\geq 0} filtration of an algebra RR, 0F1F2R0\subset F_{1}\subset F_{2}\cdots\subset R, we can get a 0\mathbb{Z}^{\geq 0}-graded algebra, the Rees algebra ReesF(R):=iFitiRees_{F}(R):=\oplus_{i}F_{i}t^{i}, where tt is a formal variable. Then Proj(ReesF(R))Proj(Rees_{F}(R)) is a compactification of Spec(R)Spec(R). When the filtration FF is fixed, we often omit FF from the notation ReesF(R)Rees_{F}(R).

Now we give the algebra191919Since we fix the dimension vector α\alpha, we omit α\alpha for the notation 𝒜\mathcal{A}. 𝒜=defHGL(V)𝒪(+)\mathcal{A}\stackrel{{\scriptstyle\mathrm{def}}}{{=}}H_{*}^{GL(V)_{\mathcal{O}}}(\mathcal{R}^{+}) a 0\mathbb{Z}_{\geq 0} filtration, which is the pullback of the filtration in remark 3.7 [BFN18] under the diagonal embedding of 00I\mathbb{Z}_{\geq 0}\xrightarrow[]{}\mathbb{Z}_{\geq 0}^{I}, as follows. Recall ni=dimVin_{i}=dimV_{i}. The GL(V)𝒪GL(V)_{\mathcal{O}}-orbits in GrGL(V)+Gr^{+}_{GL(V)} are numbered by II-colored partitions (λ(i))iI(\lambda^{(i)})_{i\in I}, λ(i)=(λ1(i)λ2(i)),\lambda^{(i)}=(\lambda^{(i)}_{1}\geq\lambda^{(i)}_{2}\geq\cdots), such that the number of parts l(λ(i))nil(\lambda^{(i)})\leq n_{i}. Given a dimension vector (mi)iI(m_{i})_{i\in I}, we define a closed GL(V)𝒪GL(V)_{\mathcal{O}}-invariant subvariety GrGL(V)¯+,(mi)GrGL(V)+\overline{Gr_{GL(V)}}^{+,(m_{i})}\subset Gr^{+}_{GL(V)} as the union of orbits GrGL(V)(λ(i))iIGr^{(\lambda^{(i)})_{i\in I}}_{GL(V)} such that λ1(i)mi\lambda^{(i)}_{1}\leq m_{i} for any iIi\in I. For mm\in\mathbb{N}, we define m++\mathcal{R}^{+}_{\leq m}\subset\mathcal{R}^{+} as the preimage of Gr¯GL(V)+,(m,,m)\overline{Gr}_{GL(V)}^{+,(m,\cdots,m)} under +GrGL(V)+\mathcal{R}^{+}\xrightarrow[]{}Gr^{+}_{GL(V)}. This gives a 0\mathbb{Z}^{\geq 0} filtration Fm=HGL(V)𝒪(m+)F_{m}=H_{*}^{GL(V)_{\mathcal{O}}}(\mathcal{R}^{+}_{\leq m}) of the algebra 𝒜.\mathcal{A}.

Define 𝐌Qα=defProj(Rees(𝒜))\mathbf{M}_{Q}^{\alpha}\stackrel{{\scriptstyle\text{def}}}{{=}}Proj(Rees(\mathcal{A})) and call 𝐌Qα\mathbf{M}_{Q}^{\alpha} the compactified Coulomb branch.

3.4 Embedding of 𝐌Qα\mathbf{M}_{Q}^{\alpha} into (HG𝒪(1+))\mathbb{P}_{\mathcal{H}}(H^{G_{\mathcal{O}}}(\mathcal{R}^{+}_{\leq 1})^{*})

We fix an ordered basis of V.V. Let TiT_{i} be corresponding diagonal subgroup of GL(Vi)GL(V_{i}) and 𝔱𝔦\mathfrak{t_{i}} be its Lie algebra. Let WiW_{i} be the Weyl group of GL(Vi).GL(V_{i}). Denote by (aji)1jni(a^{i}_{j})_{1\leq j\leq n_{i}} the corresponding standard basis of 𝔱𝔦\mathfrak{t_{i}}. Notice that in the last section, we already denote certain generators of 𝒪(Cα)\mathcal{O}(C^{\alpha}) by ajia^{i}_{j}. Here we identify HG𝒪(pt)HT𝒪(pt)H^{*}_{G_{\mathcal{O}}}(pt)\hookrightarrow H^{*}_{T_{\mathcal{O}}}(pt) with 𝒪(α)𝒪(Cα)\mathcal{O}(\mathcal{H}^{\alpha})\hookrightarrow\mathcal{O}(C^{\alpha}) via Sym((𝔱/W))Sym(𝔱)Sym((\mathfrak{t}/W)^{*})\hookrightarrow Sym(\mathfrak{t}^{*}). The algebra mHG𝒪(Rm+)tm\oplus_{m}H^{G_{\mathcal{O}}}(R^{+}_{\leq m})t^{m} is a module over HG𝒪(pt)HG(pt)H^{*}_{G_{\mathcal{O}}}(pt)\cong H^{*}_{G}(pt) so we have a projection

𝐌Qαα.\mathbf{M}_{Q}^{\alpha}\xrightarrow[]{}\mathcal{H}^{\alpha}. (18)
Lemma 6.

As a HG𝒪(pt)H^{*}_{G_{\mathcal{O}}}(pt)-algebra, the Rees algebra mHG𝒪(m+)tm\oplus_{m}H_{*}^{G_{\mathcal{O}}}(\mathcal{R}^{+}_{\leq m})t^{m} is generated by HG𝒪(1+)tH_{*}^{G_{\mathcal{O}}}(\mathcal{R}^{+}_{\leq 1})t.

Proof.

We recall some notations in [BFN19]. Let λ=π1(Grλ)\mathcal{R}_{\lambda}=\mathcal{R}\cap\pi^{-1}(Gr^{\lambda}) (λ=π1(Grλ)¯\mathcal{R}_{\leq\lambda}=\mathcal{R}\cap\pi^{-1}(\overline{Gr^{\lambda})}) be the restriction of \mathcal{R} on the G𝒪G_{\mathcal{O}}-orbit GrλGr^{\lambda} (G𝒪G_{\mathcal{O}}-orbit closure Grλ¯\overline{Gr^{\lambda}}, resp). Let <λ\mathcal{R}_{<\lambda} be the complement λλ\mathcal{R}_{\leq\lambda}\setminus\mathcal{R}_{\lambda}. It is a closed subvariety. Lemma 6.2 in [BFN18] says that the Mayer-Vietoris sequence splits into short exact sequences

0HG𝒪(<λ)HG𝒪(λ)HG𝒪(λ)0.0\xrightarrow[]{}H^{G_{\mathcal{O}}}_{*}(\mathcal{R}_{<\lambda})\xrightarrow[]{}H^{G_{\mathcal{O}}}_{*}(\mathcal{R}_{\leq\lambda})\xrightarrow[]{}H^{G_{\mathcal{O}}}_{*}(\mathcal{R}_{\lambda})\xrightarrow[]{}0.

Moreover, as HG𝒪(pt)H^{*}_{G_{\mathcal{O}}}(pt)-module, this exact sequence splits canonically so we have

HG𝒪(λ)=HG𝒪(<λ)HG𝒪(λ).H^{G_{\mathcal{O}}}_{*}(\mathcal{R}_{\leq\lambda})=H^{G_{\mathcal{O}}}_{*}(\mathcal{R}_{<\lambda})\oplus H^{G_{\mathcal{O}}}_{*}(\mathcal{R}_{\lambda}). (19)

It suffices to show that any aHG𝒪(m+)tma\in H^{G_{\mathcal{O}}}(\mathcal{R}^{+}_{\leq m})t^{m} is generated by HG𝒪(1+)tH^{G_{\mathcal{O}}}(\mathcal{R}^{+}_{\leq 1})t. Let Λ(m)\Lambda(m) be the set of all maximal λ\lambda such that λ1(i)m\lambda^{(i)}_{1}\leq m. By definition of m+\mathcal{R}^{+}_{\leq m} and Grλ¯=μ:μλGrμ\overline{Gr_{\lambda}}=\cup_{\mu:\mu\leq\lambda}Gr_{\mu}, we have m+=λΛ(m)λ+\mathcal{R}^{+}_{\leq m}=\sqcup_{\lambda\in\Lambda(m)}\mathcal{R}^{+}_{\leq\lambda}. It suffices to prove the claim for any aHG𝒪(λ+)tma\in H^{G_{\mathcal{O}}}(\mathcal{R}^{+}_{\leq\lambda})t^{m} for λΛ(m)\lambda\in\Lambda(m). We prove this by induction on λ\lambda. Suppose for any μ<λ\mu<\lambda, the theorem holds. By formula (19) and induction hypothesis, it suffices to show when aHG𝒪(λ)a\in H^{G_{\mathcal{O}}}_{*}(\mathcal{R}_{\lambda}) for some λ\lambda. The second paragraph after proposition 6.1 [BFN19] says

HG𝒪(Rλ)HStabG(λ)(pt)[λ][𝐭]Wλ[λ],H_{*}^{G_{\mathcal{O}}}(R_{\lambda})\cong H^{*}_{Stab_{G}(\lambda)}(pt)\cap[\mathcal{R}_{\lambda}]\cong\mathbb{C}[\mathbf{t}]^{W_{\lambda}}[\mathcal{R}_{\lambda}],

where \cap is the cap product, WλW_{\lambda} is the stabilizer of λ\lambda in the Weyl group WW and [λ][\mathcal{R}_{\lambda}] is the G𝒪G_{\mathcal{O}}-equivariant fundamental class of λ\mathcal{R}_{\lambda}. So it suffices to prove the case a=[λ]tm.a=[\mathcal{R}_{\lambda}]t^{m}. Recall we denote HGL(V)𝒪(+)H^{GL(V)_{\mathcal{O}}}_{*}(\mathcal{R}^{+}) by 𝒜\mathcal{A}. In the proof of [BFN19] proposition 6.8, regarding [λ][\mathcal{R}_{\lambda}] as an element in gr𝒜gr\mathcal{A}, we have

[λ][μ]=[λ+μ][\mathcal{R}_{\lambda}][\mathcal{R}_{\mu}]=[\mathcal{R}_{\lambda+\mu}]

in gr𝒜gr\mathcal{A} when λ,μ\lambda,\mu are in the same ”generalized Weyl chamber”. In this case, any weight of NN is a root of GG, so their ”generalized Weyl chamber” is the same as the usual Weyl chamber. Let ϖβ\varpi_{\beta} be the fundamental weights of GL(V)GL(V). Here, the fundamental weights of GLnGL_{n} consist of all ϖk=(1,1,1,0,0,0)\varpi_{k}=(1,1\cdots,1,0,0\cdots,0) for 1kn1\leq k\leq n where kk is the number of 11’s. The fundamental weights of the product of GL(Vi)GL(V_{i})’s consist of the disjoint union of fundamental weights of GL(Vi)GL(V_{i})’s. Since β1\mathcal{R}_{\beta}\subset\mathcal{R}_{\leq 1}, we have HG𝒪(β)tRees𝒜.H^{G_{\mathcal{O}}}(\mathcal{R}_{\beta})t\subset Rees\mathcal{A}. We can assume maxiλ1(i)=mmax_{i}\lambda^{(i)}_{1}=m. It is easy to see that we can choose mm fundamental weights βi\beta_{i} such that iβi=λ\sum_{i}\beta_{i}=\lambda. So in 𝒜\mathcal{A}, we have [β1][β2][βm][λ]HG𝒪(<λ)[\mathcal{R}_{\beta_{1}}][\mathcal{R}_{\beta_{2}}]\cdots[\mathcal{R}_{\beta_{m}}]-[\mathcal{R}_{\lambda}]\in H^{G_{\mathcal{O}}}(\mathcal{R}_{<\lambda}). In Rees𝒜Rees\mathcal{A}, we have [β1]t[β2]t[βm]t[λ]tmHG𝒪(<λ)tm[\mathcal{R}_{\beta_{1}}]t[\mathcal{R}_{\beta_{2}}]t\cdots[\mathcal{R}_{\beta_{m}}]t-[\mathcal{R}_{\lambda}]t^{m}\in H_{*}^{G_{\mathcal{O}}}(\mathcal{R}_{<\lambda})t^{m}, which is generated by HG𝒪(1)H_{*}^{G_{\mathcal{O}}}(\mathcal{R}_{\leq 1}) by induction. Hence we claim [λ]tm[\mathcal{R}_{\lambda}]t^{m} is generated by HG𝒪(1)H_{*}^{G_{\mathcal{O}}}(\mathcal{R}_{\leq 1}).

Corollary 1.

We have an embedding

𝐌αjαα(HG(1+)),\mathbf{M}^{\alpha}\xhookrightarrow{j_{\alpha}}\mathbb{P}_{\mathcal{H}^{\alpha}}(H_{*}^{G}(\mathcal{R}^{+}_{\leq 1})^{*}),

where HG(1+)H_{*}^{G}(\mathcal{R}^{+}_{\leq 1})^{*} is understood as a vector bundle over Spec(HG(pt))α.Spec(H^{*}_{G}(pt))\cong\mathcal{H}^{\alpha}.

Proof.

By lemma 6, we have the surjection of graded algebras

𝒮(HG𝒪(1+)t)mHG𝒪(m+)tm.\mathcal{S}(H_{*}^{G_{\mathcal{O}}}(\mathcal{R}^{+}_{\leq 1})t)\twoheadrightarrow\oplus_{m}H_{*}^{G_{\mathcal{O}}}(\mathcal{R}^{+}_{\leq m})t^{m}.

Taking Proj, we get the embedding. ∎

3.5 Bases of HG𝒪(1+)H^{G_{\mathcal{O}}}_{*}(\mathcal{R}^{+}_{\leq 1}) as HG𝒪(pt)𝒪(α)H_{G_{\mathcal{O}}}^{*}(pt)\cong\mathcal{O}(\mathcal{H}^{\alpha})-module

Lemma 7.
HG𝒪(1+)(ki)iIiIHGL(Vi)(Gr(ni,ki)).H^{G_{\mathcal{O}}}_{*}(\mathcal{R}^{+}_{\leq 1})\cong\bigoplus_{(k_{i})_{i\in I}}\bigotimes_{i\in I}H^{GL(V_{i})}_{*}(Gr(n_{i},k_{i})). (20)
Proof.

Let {ϖkii,1kini}\{\varpi^{i}_{k_{i}},1\leq k_{i}\leq n_{i}\} be all fundamental weights of GL(Vi)GL(V_{i}), where ϖkii=(1,1,1,0,0,0)\varpi^{i}_{k_{i}}=(1,1\cdots,1,0,0\cdots,0), kik_{i}=number of 11’s. Now we allow ki=0k_{i}=0, denote ϖ0i=(0,,0)\varpi^{i}_{0}=(0,\cdots,0) and let ϖ(ki)iI=def(ϖkii)iI\varpi_{(k_{i})_{i\in I}}\stackrel{{\scriptstyle def}}{{=}}(\varpi^{i}_{k_{i}})_{i\in I}. Recall V=iIViV=\oplus_{i\in I}V_{i} with dimension vector α=(ni)iI.\alpha=(n_{i})_{i\in I}. For the set {ϖ(ki)iI,0kini,iI}\{\varpi_{(k_{i})_{i\in I}},0\leq k_{i}\leq n_{i},i\in I\}, we have

1+=0kini,iIϖ(ki)iI.\mathcal{R}^{+}_{\leq 1}=\bigsqcup_{0\leq k_{i}\leq n_{i},i\in I}\mathcal{R}_{\varpi_{(k_{i})_{i\in I}}}.

For each component of 1+\mathcal{R}^{+}_{\leq 1}, ϖ(ki)iI\mathcal{R}_{\varpi_{(k_{i})_{i\in I}}} is a vector bundle over the GL(V)𝒪GL(V)_{\mathcal{O}}-orbit 𝒢ϖ(ki)iI=iGr(ni,ki)\mathcal{G}_{\varpi_{(k_{i})_{i\in I}}}=\prod_{i}Gr(n_{i},k_{i}) so their homologies are isomorphic (without grading). ∎

Now we fix a dimension vector β=(ki)iI\beta=(k_{i})_{i\in I}. By Poincare duality, HG(ϖβ)𝑝HG(ϖβ)H_{G}^{*}(\mathcal{R}_{\varpi_{\beta}})\xrightarrow[\sim]{p}H^{G}_{*}(\mathcal{R}_{\varpi_{\beta}}), where202020again we ignore the grading p(c)=c[ϖβ]Gp(c)=c\cap[\mathcal{R}_{\varpi_{\beta}}]^{G} and [ϖβ]G[\mathcal{R}_{\varpi_{\beta}}]^{G} is the GG-equivariant fundamental class of ϖβ\mathcal{R}_{\varpi_{\beta}}. The homology group HG(ϖβ)H^{G}_{*}(\mathcal{R}_{\varpi_{\beta}}) is a free rank 1 module over the ring HG(iIGr(ni,ki))iIHGL(Vi)Gr(ni,ki).H_{G}^{*}(\prod_{i\in I}Gr(n_{i},k_{i}))\cong\otimes_{i\in I}H_{GL(V_{i})}^{*}Gr(n_{i},k_{i}). For the cohomology of Grassmannian, we have a well-known result. Since the cohomology of the product of Grassmannians is just the tensor product of cohomologies of Grassmannians, for brevity, in lemma 8, we set I={1}I=\{1\} so GL(V)=GLnGL(V)=GL_{n} and we abbreviate aj1a^{1}_{j} by aja_{j} for 1jn.1\leq j\leq n. Let SS be the tautological bundle over Gr(n,k)Gr(n,k) and QQ be the quotient bundle. Denote cic_{i} the ii-th GG-equivariant Chern class and cc the total Chern class. Recall that the notation fsf^{s} is the elementary symmetric functions, introduced before theorem 1.

Lemma 8.
HGL(V)Gr(n,k)HGL(V)(pt)[cl(S),cj(Q)]/I,H_{GL(V)}^{*}Gr(n,k)\cong H_{GL(V)}^{*}(pt)[c_{l}(S),c_{j}(Q)]/\mathrm{I},

where the ideal I\mathrm{I} is generated by

1lk,l+j=scl(S)cj(Q)fs(a1,,an)\sum_{1\leq l\leq k,l+j=s}c_{l}(S)c_{j}(Q)-f^{s}(a_{1},\cdots,a_{n})

for all ss such that 1sn.{1\leq s\leq n}.

Proof.

We have an exact sequence 0SVQ00\xrightarrow[]{}S\xrightarrow[]{}V\xrightarrow[]{}Q\xrightarrow[]{}0. It is well-known that as an HGL(V)(pt)H_{GL(V)}^{*}(pt)-algebra, HT(Gr(n,k)H^{*}_{T}(Gr(n,k) has generators c1(S),,ck(S),c1(Q),cnk(Q)c_{1}(S),\cdots,c_{k}(S),c_{1}(Q),\cdots c_{n-k}(Q). From the exact sequence, we have c(S)c(Q)=c(V)c(S)c(Q)=c(V). c(V)=χc(Vχ)=χ(1+χ)=i(1+ai)c(V)=\prod_{\chi}c(V_{\chi})=\prod_{\chi}(1+\chi)=\prod_{i}(1+a_{i}), where χ\chi are characters of TT and VχV_{\chi} is the χ\chi-weight space of VV. Plug in c=icic=\sum_{i}c^{i} into S,QS,Q and expand c(S)c(Q)=c(V)c(S)c(Q)=c(V). Comparing the degree ss part, we get the relation in the lemma.

Remark 5.

By the same argument, we get the isomorphism for TT-equivariant cohomology.

HT(Gr(n,k))HT(pt)[clT(S),cjT(Q)]/I,H_{T}^{*}(Gr(n,k))\cong H_{T}^{*}(pt)[c_{l}^{T}(S),c_{j}^{T}(Q)]/I,

for the ideal II with the same generators from the lemma.

3.6 Pulling back 𝐌α\mathbf{M}^{\alpha} under CαqααC^{\alpha}\xrightarrow[]{q^{\alpha}}\mathcal{H}^{\alpha}

To apply localization theory in (3.1), we consider HT𝒪()H^{T_{\mathcal{O}}}_{*}(\mathcal{R}) which is an algebra over HT𝒪(pt)H_{*}^{T_{\mathcal{O}}}(pt). We identify Spec(HT𝒪(pt))Spec(H_{*}^{T_{\mathcal{O}}}(pt)) with CαC^{\alpha}.

Lemma 9.

The pullback of Spec(HG𝒪())Spec(H^{G_{\mathcal{O}}}_{*}(\mathcal{R})), Mα\textsf{M}^{\alpha} and 𝐌α\mathbf{M}^{\alpha} under the quotient map CαqααC^{\alpha}\xrightarrow[]{q^{\alpha}}\mathcal{H}^{\alpha} is Spec(HT𝒪())Spec(H^{T_{\mathcal{O}}}_{*}(\mathcal{R})), Spec(HT𝒪(+))Spec(H^{T_{\mathcal{O}}}_{*}(\mathcal{R}^{+})) and Proj(mHT𝒪(m+)tm)Proj(\oplus_{m}H_{*}^{T_{\mathcal{O}}}(\mathcal{R}^{+}_{\leq m})t^{m})

Proof.

The following is a Cartesian diagram HT𝒪(?){H^{T_{\mathcal{O}}}_{*}(?)}HG𝒪(?){H^{G_{\mathcal{O}}}_{*}(?)}HT𝒪(pt){H^{T_{\mathcal{O}}}_{*}(pt)}HG𝒪(pt){H^{G_{\mathcal{O}}}_{*}(pt)} where ?? is \mathcal{R}, +\mathcal{R}^{+} and m+\mathcal{R}_{\leq m}^{+}. ∎

Lemma 10.

For the quiver gauge theory case, under the identification of 𝔱\mathfrak{t} and CαC^{\alpha}, the pushforward homomorphism ι\iota_{*} for the inclusion T,NT𝜄\mathcal{R}_{T,N_{T}}\xrightarrow[]{\iota}\mathcal{R}

HT𝒪(T,NT)ιHT𝒪()H^{T_{\mathcal{O}}}_{*}(\mathcal{R}_{T,N_{T}})\xrightarrow[]{\iota_{*}}H^{T_{\mathcal{O}}}_{*}(\mathcal{R})

becomes an isomorphism over CregαC^{\alpha}_{reg}.

Proof.

In (16), it says the pushforward homomorphism ι\iota_{*} becomes an isomorphism over 𝔱reg\mathfrak{t}_{reg}. For each direct summand Hom(Vi,Vi)Hom(V_{i},V_{i^{\prime}}) in NN, the localization inverts all aliajia^{i}_{l}-a^{i^{\prime}}_{j} for 1ldimVi,1jdimVi1\leq l\leq dimV_{i},1\leq j\leq dimV_{i^{\prime}}. It is clear that the localization of CαC^{\alpha} to CregαC^{\alpha}_{reg} includes all roots of GG and weights of NTN_{T}. ∎

We define T,NT+\mathcal{R}^{+}_{T,N_{T}} and its filtration (T,NT+)m(\mathcal{R}^{+}_{T,N_{T}})_{\leq m} as the pullback under ι\iota of +\mathcal{R}^{+} and its filtration m+\mathcal{R}_{\leq m}^{+}. Then we get the Rees algebra mHT𝒪((T,NT+)m)tm\oplus_{m}H_{*}^{T_{\mathcal{O}}}((\mathcal{R}^{+}_{T,N_{T}})_{\leq m})t^{m} of HT𝒪(T,NT).H^{T_{\mathcal{O}}}_{*}(\mathcal{R}_{T,N_{T}}). For any space XX, we denote

T(X)=defHT(X)𝒪(Cα)𝒪(Cα)reg.\mathcal{H}^{T}(X)\stackrel{{\scriptstyle def}}{{=}}H^{T}(X)\otimes_{\mathcal{O}(C^{\alpha})}\mathcal{O}(C^{\alpha})_{reg}. (21)

For example, by this convention, the localized Rees algebra of T,NT+\mathcal{R}^{+}_{T,N_{T}} is denoted by mT𝒪((T,NT)m+)tm,\oplus_{m}\mathcal{H}^{T_{\mathcal{O}}}((\mathcal{R}_{T,N_{T}})^{+}_{\leq m})t^{m},

We fix some notations. In our case 𝒢(T)\mathcal{G}(T) is the disjoint union of points tχ=(ti)iIχit^{\chi}=(t_{i})^{\chi^{i}}_{i\in I} where χ=(χi)iI\chi=(\chi^{i})_{i\in I} is a cocharacter of T=iITiT=\prod_{i\in I}T_{i} and χi\chi^{i} is a cocharactor of TiT_{i}. By the standard isomorphism Ti(Gm)niT_{i}\cong(G_{m})^{n_{i}} we can write χi=(χji)1jni\chi^{i}=(\chi_{j}^{i})_{1\leq j\leq n_{i}}.

3.6.1 Generators of the localized Rees algebra mT𝒪((T,NT)m+)tm\oplus_{m}\mathcal{H}^{T_{\mathcal{O}}}((\mathcal{R}_{T,N_{T}})^{+}_{\leq m})t^{m}

From the definition of the filtration of T,NT\mathcal{R}_{T,N_{T}} , it is easy to see that

(T,NT)m+=0χjim for any i,jVχ.(\mathcal{R}_{T,N_{T}})^{+}_{\leq m}=\bigsqcup_{0\leq\chi^{i}_{j}\leq m\text{ for any }i,j}V_{\chi}.
Lemma 11.

The localized Rees algebra mT𝒪((T,NT)m+)tm\oplus_{m}\mathcal{H}^{T_{\mathcal{O}}}((\mathcal{R}_{T,N_{T}})^{+}_{\leq m})t^{m} is generated by T𝒪((T,NT)1+)t.\mathcal{H}^{T_{\mathcal{O}}}((\mathcal{R}_{T,N_{T}})^{+}_{\leq 1})t.

Proof.

This follows from the formula (17) in theorem 2. ∎

3.7 HT𝒪((T,NT+)1)ιHT𝒪(1+)H^{T_{\mathcal{O}}}_{*}((\mathcal{R}_{T,N_{T}}^{+})_{\leq 1})\xrightarrow[]{\iota_{*}}H^{T_{\mathcal{O}}}_{*}(\mathcal{R}^{+}_{\leq 1}) over CregC_{reg}

In the next lemma, we will study the map

HT𝒪(1+)T𝒪(1+)(ι)1T𝒪(T,NT)1+)χ|0χji1 for any i,jT𝒪(Vχ).H^{T_{\mathcal{O}}}_{*}(\mathcal{R}^{+}_{\leq 1})\xhookrightarrow{}\mathcal{H}^{T_{\mathcal{O}}}_{*}(\mathcal{R}^{+}_{\leq 1})\xrightarrow[]{(\iota_{*})^{-1}}\mathcal{H}^{T_{\mathcal{O}}}_{*}(\mathcal{R}_{T,N_{T}})^{+}_{\leq 1})\cong\bigoplus_{\chi|0\leq\chi^{i}_{j}\leq 1\text{ for any }i,j}\mathcal{H}^{T_{\mathcal{O}}}(V_{\chi}).

Since the preimage of ϖβ\mathcal{R}_{\varpi_{\beta}} under ι\iota is χWϖβVχ\sqcup_{\chi\in W\varpi_{\beta}}V_{\chi}, restricting the above map to HT𝒪(ϖβ)H_{*}^{T_{\mathcal{O}}}(\mathcal{R}_{\varpi_{\beta}}) we get

HT𝒪(ϖβ)χWϖβHT𝒪(Vχ).H^{T_{\mathcal{O}}}_{*}(\mathcal{R}_{\varpi_{\beta}})\xhookrightarrow{}\bigoplus_{\chi\in W\varpi_{\beta}}H_{*}^{T_{\mathcal{O}}}(V_{\chi}). (22)

For each χWϖβ\chi\in W\varpi_{\beta}, let xχx^{\chi} be the χ\chi component of the image under (22) of the fundamental class of [ϖβ][\mathcal{R}_{\varpi_{\beta}}]. For any χ\chi such that 0χji10\leq\chi^{i}_{j}\leq 1 for all i,ji,j, we define a set S(χ)=iIS(χi)S(\chi)=\sqcup_{i\in I}S(\chi_{i}) where S(χi)={j|χji=1}.S(\chi_{i})=\{j|\chi^{i}_{j}=1\}. It is clear such χ\chi is in bijection with SS such that SSα,|S|=βS\subset S^{\alpha},|S|=\beta. So given SS, we also use the notation χ(S)\chi(S), meaning the χ\chi such that S(χ)S(\chi) is the given set SS. We denote the set {1,,ni}\{1,\cdots,n_{i}\} by [ni][n_{i}]. Recall we denote the fundamental class [Vχ][V_{\chi}] by rχHT𝒪(Vχ)r^{\chi}\in H_{*}^{T_{\mathcal{O}}}(V_{\chi}).

Lemma 12.

a)

xχ=(Euχ)1rχ,x^{\chi}=(Eu_{\chi})^{-1}r^{\chi},

where EuχEu_{\chi} is the TT-equivariant Euler class of Nϖβ/VχN_{\mathcal{R}_{\varpi_{\beta}}/V_{\chi}} at the point tχt^{\chi}. Here Nϖβ/VχN_{\mathcal{R}_{\varpi_{\beta}}/V_{\chi}} is the normal bundle of VχV_{\chi} in ϖβ.\mathcal{R}_{\varpi_{\beta}}.
b)

Euχ=iIjS(χi),l[ni]S(χi)(ajiali).Eu_{\chi}=\prod_{i\in I}\prod_{j\in S(\chi_{i}),l\in[n_{i}]\setminus S(\chi_{i})}(a^{i}_{j}-a^{i}_{l}).
Proof.

a) This is just localization theory. b) Let (eji)iI,j[ni](e^{i}_{j})_{i\in I,j\in[n_{i}]} be the TT-eigenvectors of V=iIVi.V=\oplus_{i\in I}V_{i}. Since the fiber of the projection ϖβ𝒢(GL(V))ϖβ=iIGr(ni,ki))\mathcal{R}_{\varpi_{\beta}}\xrightarrow[]{}\mathcal{G}(GL(V))_{\varpi_{\beta}}=\prod_{i\in I}Gr(n_{i},k_{i})) at tχt^{\chi} is VχV_{\chi}, we have

(Nϖβ/Vχ)tχ=T[iIjS(χi)keji](iIGr(ni,ki)),(N_{\mathcal{R}_{\varpi_{\beta}}/V_{\chi}})_{t^{\chi}}=T_{[\oplus_{i\in I}\oplus_{j\in S(\chi_{i})}ke^{i}_{j}]}(\prod_{i\in I}Gr(n_{i},k_{i})),

where [iIjS(χi)keji][\oplus_{i\in I}\oplus_{j\in S(\chi_{i})}ke^{i}_{j}] is the point tχt^{\chi} viewed as a point in iIGr(ni,ki)\prod_{i\in I}Gr(n_{i},k_{i}) (i.e., a vector subspace in VV). Now the formula follows from a standard computation about weight spaces of tangent spaces of a partial flag variety. ∎

Corollary 2.

The localized Rees algebra mT𝒪((T,NT)m+)tm\oplus_{m}\mathcal{H}^{T_{\mathcal{O}}}((\mathcal{R}_{T,N_{T}})^{+}_{\leq m})t^{m} is generated by xχ(S)t,SSαx^{\chi(S)}t,S\subset S^{\alpha}.

Proof.

Follows from lemma 11 and lemma 12 part (a). ∎

4 Identification of 𝐌α\mathbf{M}^{\alpha} and ZαZ^{\alpha}

For a quiver QQ, we forget the direction of arrows and define an integer matrix κ(Q)\kappa(Q) as κii=1\kappa_{ii}=1- number of self-loop of ii and κij\kappa_{ij}=-number of arrow between ii and jj. For a symmetric integral matrix κ\kappa, if κii1\kappa_{ii}\leq 1 for all iIi\in I and κij0\kappa_{ij}\leq 0 for all i,jI,iji,j\in I,i\neq j, we say κ\kappa is of quiver-type, since in this case there exists a quiver QQ such that κ=κ(Q).\kappa=\kappa(Q).

Remark 6 (Different conventions about the matrix κ(Q)\kappa(Q)).

For a diagram QQ (possibly with self-loops), define the Cartan matrix C(Q)C(Q) associated with QQ as Cii=22C_{ii}=2-2 number of self-loop of ii and CijC_{ij}=-number of arrow between ii and jj. So the relation is Cii=2κii,Cij=κij.C_{ii}=2\kappa_{ii},C_{ij}=\kappa_{ij}. For QQ of finite ADE type, let GQG_{Q} be the simply-connected simple algebraic group corresponding to QQ. As mentioned in the introduction, in [MYZ21, Mir23], Mirkovic proved that Zκ(Q)αZ^{\alpha}_{\kappa(Q)}212121In our notation is isomorphic to the Zastava space ZGQαZ^{\alpha}_{G_{Q}}. The notation κ\kappa in [MYZ21, Mir23] is C(Q)C(Q) here.

Theorem 3.

For a quiver QQ, the Compactified Coulomb branch 𝐌Qα\mathbf{M}_{Q}^{\alpha} is canonically isomorphic to the local projective space Zκ(Q)α.Z_{\kappa(Q)}^{\alpha}.

Again, we will prove it after pullback under Cα𝑎αC^{\alpha}\xrightarrow[]{a}\mathcal{H}^{\alpha}. We list what we have.

(qα)𝐌α(q^{\alpha})^{*}\mathbf{M}^{\alpha} (qα)Zα(q^{\alpha})^{*}Z^{\alpha}
embedding (qα)𝐌αjαCα(HT(1+))(q^{\alpha})^{*}\mathbf{M}^{\alpha}\xhookrightarrow{j_{\alpha}}\mathbb{P}_{C^{\alpha}}(H_{*}^{T}(\mathcal{R}^{+}_{\leq 1})^{*}) (a) (qα)ZαiαCα((qα)Vα)(q^{\alpha})^{*}Z^{\alpha}\xhookrightarrow{i_{\alpha}}\mathbb{P}_{C^{\alpha}}((q^{\alpha})^{*}V^{\alpha}) (b)
over regular part CregαC^{\alpha}_{reg} product of 1\mathbb{P}^{1} (c) product of 1\mathbb{P}^{1} (d)
global basis basis of HT(1+)H_{*}^{T}(\mathcal{R}^{+}_{\leq 1}) (e) basis of Γ(Cα,(qα)α)\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha}) (f)
local basis sSs^{S} (section 2.7.3) xχx^{\chi} (section 3.7, after (22))

(a) is by Corollary 1. (b) is by section 2.5. (c) is not proven yet but is easy to see. It is listed as a heuristic for the proof of theorem 3. (d) is by definition. (e) is by section 3.5. (f) is by section 2.7.2.

4.1 Identification of HT(1+)H_{*}^{T}(\mathcal{R}^{+}_{\leq 1}) and Γ(Cα,(qα))\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I})

We have the following decompositions into connected components.

1\mathcal{R}_{\leq 1} (qα)Gr(𝒯α)(q^{\alpha})^{*}Gr(\mathcal{T}^{\alpha})
connected components ϖβ\mathcal{R}_{\varpi_{\beta}} (qα)Grβ(𝒯α)(q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha})

The connected components on both sides are indexed by the same data β=(ki)iI\beta=(k_{i})_{i\in I} and 0kini0\leq k_{i}\leq n_{i}. On each component, we have

rank 1 free module HT(ϖβ)H_{*}^{T}(\mathcal{R}_{\varpi_{\beta}}) Γ(Cα,(qα)α,β)\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta})
over the ring HT(ϖβ)H^{*}_{T}(\mathcal{R}_{\varpi_{\beta}}) 𝒪((qα)Grβ(𝒯α))\mathcal{O}((q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha}))
with basis the fundamental class [ϖβ][\mathcal{R}_{\varpi_{\beta}}] uβ1u^{\beta}\otimes 1

In the third line, the fundamental class [ϖβ][\mathcal{R}_{\varpi_{\beta}}] is a basis of HT(ϖβ)H_{*}^{T}(\mathcal{R}_{\varpi_{\beta}}) over HT(ϖβ)H^{*}_{T}(\mathcal{R}_{\varpi_{\beta}}) by Poincare duality (See the paragraph after lemma 7, note there we used GG-equivariance and here TT-equivariance). The basis of Γ(Cα,(qα)α,β)\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta}) over 𝒪((qα)Grβ(𝒯α))\mathcal{O}((q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha})) is chosen as uβ1u^{\beta}\otimes 1 (defined at the beginning of 2.7.3222222There, it is considered as a basis element for Γ(Cα,(qα)α,β)\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta}) as an 𝒪(Cα)\mathcal{O}(C^{\alpha})-module. Here it is a basis as an 𝒪((qα)Grβ(𝒯α,β))\mathcal{O}((q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha,\beta}))-module.).

We will identify the rings in lemma 13 and therefore also the trivialized modules from line 1 of the above table.

Lemma 13.

There are canonical ring isomorphisms of the horizontal arrows of the following diagram

𝒪((qα)Grβ(𝒯α)){\mathcal{O}((q^{\alpha})^{*}Gr^{\beta}(\mathcal{T}^{\alpha}))}HT(ϖβ)HT(iIGr(ni,ki)){H^{*}_{T}(\mathcal{R}_{\varpi_{\beta}})\cong H^{*}_{T}(\prod_{i\in I}Gr(n_{i},k_{i}))}S,|S|=|β|𝒪(Cregα){\bigoplus_{S,|S|=|\beta|}\mathcal{O}(C^{\alpha}_{reg})}χWϖβT(tχ){\bigoplus_{\chi\in W\varpi_{\beta}}\mathcal{H}^{*}_{T}(t^{\chi})}\scriptstyle{\cong}res\scriptstyle{res}loc\scriptstyle{loc}\scriptstyle{\cong}

which makes the diagram commute. Here locloc is the localization map.

Proof.

The isomorphism for the upper horizontal map is clear from theorem 1 and remark 5, where in the case I={1}I=\{1\}, it sends clic^{i}_{l} to Ci(Sl)C_{i}(S_{l}) and dkid^{i}_{k} to Ci(Qk)C_{i}(Q_{k}). The definition of the lower horizontal map and the commutativity of the diagram follow from the localization theory of equivariant homology. ∎

Denote the above isomorphism between these two rank 1 modules (by the above map between rings and trivialization maps from the bases [ϖβ][\mathcal{R}_{\varpi_{\beta}}] and uβ1u^{\beta}\otimes 1) by relrel,

Γ(Cα,(qα)α,β)relHT(ϖβ), rel(uβ1)=[ϖβ].\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta})\xrightarrow[\cong]{rel}H_{*}^{T}(\mathcal{R}_{\varpi_{\beta}}),\text{ }rel(u^{\beta}\otimes 1)=[\mathcal{R}_{\varpi_{\beta}}]. (23)

4.2 Matching the local basis

We recall a lemma in the localization theory of equivariant homology theory.

Lemma 14.

Let a torus TT act on a complex algebraic variety XX. The action of HT(X)H^{*}_{T}(X) on HT(X)H_{*}^{T}(X) is compatible with localizations, i.e.

HT(X)×HT(X){H^{*}_{T}(X)\times H_{*}^{T}(X)}HT(X){H_{*}^{T}(X)}HT(XT)×HT(XT){H^{*}_{T}(X^{T})\times H_{*}^{T}(X^{T})}HT(XT){H_{*}^{T}(X^{T})}HT(XT)×T(XT){H^{*}_{T}(X^{T})\times\mathcal{H}_{*}^{T}(X^{T})}T(XT).{\mathcal{H}_{*}^{T}(X^{T}).}loc×loc\scriptstyle{loc\times loc}loc\scriptstyle{loc}

Now we state a lemma about the decomposition of the two modules in 23 after localizations.

Lemma 15.

The following diagram commutes
Γ(Cα,(qα)α,β){\Gamma(C^{\alpha},(q^{\alpha})^{*}\mathcal{I}^{\alpha,\beta})}HT(ϖβ){H_{*}^{T}(\mathcal{R}_{\varpi_{\beta}})}S,|S|=βΓ(Cregα,Lregα,S){\bigoplus_{S,|S|=\beta}\Gamma(C_{reg}^{\alpha},L^{\alpha,S}_{reg})}χWϖβT(tχ),{\bigoplus_{\chi\in W\varpi_{\beta}}\mathcal{H}_{*}^{T}(t^{\chi}),}rel\scriptstyle{rel}\scriptstyle{\cong}loc\scriptstyle{loc}loc\scriptstyle{loc}rel\scriptstyle{rel}\scriptstyle{\cong}
where each term is considered as a module over the ring corresponding to the same place in the diagram in lemma 13. Moreover, when we decompose the terms on the bottom as 𝒪(Cregα)\mathcal{O}(C^{\alpha}_{reg})(resp. T(pt)\mathcal{H}^{T}_{*}(pt))-modules, the decomposition is compatible with relrel, i.e. it maps the SS-summand to the χ(S)\chi(S)-summand.

Proof.

Follows from lemmas 5 and 14. ∎

Lemma 16.

Restricting relrel to CregαC_{reg}^{\alpha}, the basis sSs^{S} (section 2.7.3) of Γ(Cregα,Lregα,S)\Gamma(C_{reg}^{\alpha},L^{\alpha,S}_{reg}) goes to the basis xχ(S)x^{\chi(S)} (section 3.7, after (22)) of T(tχ)\mathcal{H}_{*}^{T}(t^{\chi}).

Proof.

From

rel(uβ1)=[ϖβ],rel(u^{\beta}\otimes 1)=[\mathcal{R}_{\varpi_{\beta}}],

decompose uβ1u^{\beta}\otimes 1 and [ϖβ][\mathcal{R}_{\varpi_{\beta}}] into SS (or χ(S)\chi(S) resp.)-component. Since sSs^{S} and xχ(S)x^{\chi(S)} are the SS-component of uβ1u^{\beta}\otimes 1 and [ϖβ][\mathcal{R}_{\varpi_{\beta}}] respectively, by lemma 15,

rel(sS)=xχ(S).rel(s^{S})=x^{\chi(S)}.

4.3 A flatness lemma

Lemma 17.

Let SS be a base space and USU\subset S an open dense subspace. Let Y=SnY=\mathbb{P}^{n}_{S} over SS. Let XX over UU be a closed subscheme in YU=UnY_{U}=\mathbb{P}^{n}_{U} and X¯\overline{X} be its closure in Y.Y. Let 𝒳\mathcal{X} be a closed subscheme flat over SS. If 𝒳|U=X\mathcal{X}|_{U}=X, we have 𝒳=X¯.\mathcal{X}=\overline{X}.

Proof.

Since 𝒳\mathcal{X} is closed in YY, X¯𝒳\overline{X}\subset\mathcal{X}. So for nn big enough, we have

𝒪(n)𝒳/S𝒪(n)X¯/S.\mathcal{O}(n)_{\mathcal{X}/S}\twoheadrightarrow\mathcal{O}(n)_{\overline{X}/S}.

Since 𝒳\mathcal{X} is flat over SS, the sheaf 𝒪(n)𝒳/S\mathcal{O}(n)_{\mathcal{X}/S} is locally free for nn big enough. Now we claim the composition

𝒪(n)𝒳/S𝒪(n)X¯/S𝒪(n)X/S\mathcal{O}(n)_{\mathcal{X}/S}\xrightarrow[]{}\mathcal{O}(n)_{\overline{X}/S}\xrightarrow[]{}\mathcal{O}(n)_{X/S}

is injective. Take a section ss in the kernel. By the assumption 𝒳|U=X\mathcal{X}|_{U}=X, s|U=0s|U=0 but ss is a section in the locally free sheaf so it must be 0. Hence

𝒪(n)𝒳/S𝒪(n)X¯/S\mathcal{O}(n)_{\mathcal{X}/S}\xrightarrow[]{}\mathcal{O}(n)_{\overline{X}/S}

is also injective so it is a isomorphism for n>>0n>>0. This implies 𝒳=X¯.\mathcal{X}=\overline{X}.

4.4 Theorem 3 reduces to comparison on the regular part

Lemma 18.

The dotted arrow exists as an isomorphism, i.e. the isomorphism relrel restricts to an isomorphism for the two embeddings.

(qα)Zregα{(q^{\alpha})^{*}Z_{reg}^{\alpha}}(qα)Zα{(q^{\alpha})^{*}Z^{\alpha}}Cα((qα)Vα){\mathbb{P}_{C^{\alpha}}((q^{\alpha})^{*}V^{\alpha})}(qα)𝐌regα{(q^{\alpha})^{*}\mathbf{M}_{reg}^{\alpha}}(qα)𝐌α{(q^{\alpha})^{*}\mathbf{M}^{\alpha}}Cα(HT(1+)){\mathbb{P}_{C^{\alpha}}(H_{*}^{T}(\mathcal{R}^{+}_{\leq 1})^{*})}iα\scriptstyle{i_{\alpha}}rel\scriptstyle{\simeq}jα\scriptstyle{j_{\alpha}}

Given this lemma, we can prove theorem 3.

Proof of theorem 3.

By [BFN18, lemma 5.3], (qα)𝐌α(q^{\alpha})^{*}\mathbf{M}^{\alpha} is flat over Cα.C^{\alpha}. Now we apply lemma 17. Let S=CαS=C^{\alpha} and U=Cregα.U=C^{\alpha}_{reg}. Let X=relreg((qα)Zregα)X=rel_{reg}((q^{\alpha})^{*}Z^{\alpha}_{reg}), 𝒳=(qα)𝐌α\mathcal{X}=(q^{\alpha})^{*}\mathbf{M}^{\alpha} and Y=Cα(HT(1+)).Y=\mathbb{P}_{C^{\alpha}}(H_{*}^{T}(\mathcal{R}^{+}_{\leq 1})^{*}). By lemma 18, we have 𝒳|U=X\mathcal{X}|_{U}=X so applying lemma 17 gives 𝒳=X\mathcal{X}=X, i.e., rel((qα)Z)=(qα)𝐌α.rel((q^{\alpha})^{*}Z)=(q^{\alpha})^{*}\mathbf{M}^{\alpha}.

Corollary 3.

The local space Zκ(Q)αZ_{\kappa(Q)}^{\alpha} is flat over α\mathcal{H}^{\alpha}.

Proof.

The space (qα)𝐌α(q^{\alpha})^{*}\mathbf{M}^{\alpha} is flat over CαC^{\alpha} (see the proof of theorem 3), so 𝐌α\mathbf{M}^{\alpha} is flat over α\mathcal{H}^{\alpha}. By theorem 3, Zκ(Q)α𝐌αZ_{\kappa(Q)}^{\alpha}\cong\mathbf{M}^{\alpha} is flat over α\mathcal{H}^{\alpha}. ∎

4.5 Proof of the comparison lemma on the regular part (lemma 18)

Proof.

Recall that sS,SSαs^{S},S\subset S^{\alpha} generate the algebra A0,regA_{0,reg} (section 2.7.3) and xχ(S)t,SSαx^{\chi(S)}t,S\subset S^{\alpha} generate the localized Rees algebra mT𝒪((T,NT)m+)tm\oplus_{m}\mathcal{H}^{T_{\mathcal{O}}}((\mathcal{R}_{T,N_{T}})^{+}_{\leq m})t^{m} (Corollary 2). It suffices to prove that these satisfy the same relations when we identify sSs^{S} with xχ(S)x^{\chi(S)} by lemma 16.

By the multiplication formula (17) in theorem 2, we get the multiplication formula in the Rees algebra, (we need to add tt on the right hand side)

rλrμ=i=1nξid(ξi(λ),ξi(μ))rλ+μt.r^{\lambda}r^{\mu}=\prod^{n}_{i=1}\xi_{i}^{d(\xi_{i}(\lambda),\xi_{i}(\mu))}r^{\lambda+\mu}t.

Replace the indices λ,μ\lambda,\mu by A,BA,B and denote the coefficient before rλ+μr^{\lambda+\mu} by fcQ(A,B)\text{fc}_{Q}(A,B), we rewrite the formula as

rArB=fcQ(A,B)rλ+μt.r^{A}r^{B}=\text{fc}_{Q}(A,B)r^{\lambda+\mu}t.

In particular,

rABrAB=fcQ(AB,AB)rλ+μt.r^{A\cup B}r^{A\cap B}=\text{fc}_{Q}(A\cup B,A\cap B)r^{\lambda+\mu}t.

hence the relations in the localized Rees algebra mT𝒪((T,NT)m+)tm\oplus_{m}\mathcal{H}^{T_{\mathcal{O}}}((\mathcal{R}_{T,N_{T}})^{+}_{\leq m})t^{m} can be written as

rArB=fcQ(A,B)fcQ(AB,AB)rABrAB.r^{A}r^{B}=\frac{\text{fc}_{Q}(A,B)}{\text{fc}_{Q}(A\cup B,A\cap B)}r^{A\cup B}r^{A\cap B}.

Recall by lemma 12 that

xS=(EuS)1rSx^{S}=(Eu_{S})^{-1}r^{S}

so the relations in terms of xSx^{S} are

xAxB=fcQ(A,B)fcQ(AB,AB)Eu(AB)Eu(AB)Eu(A)Eu(B)xABxAB.x^{A}x^{B}=\frac{\text{fc}_{Q}(A,B)}{\text{fc}_{Q}(A\cup B,A\cap B)}\frac{Eu(A\cup B)Eu(A\cap B)}{Eu(A)Eu(B)}x^{A\cup B}x^{A\cap B}.

Now we consider the relations in the algebra A0,reg.A_{0,reg}. Recall in (12), we defined l(β)l(\beta) depending on κ.\kappa. Now we set sβs^{\beta} by the same formula as in (11) replacing l(β)l(\beta) by

lQ(β)=aE such that s(a)=i,t(a)=i,iil(Sβ)i,j(Sβ)i(xlixji)κiiiI,lj,l,j(Sβ)i(xlixji)κii,l_{Q}(\beta)=\prod_{a\in E\text{ such that }s(a)=i,t(a)=i^{\prime},i\neq i^{\prime}l\in(S_{\beta})_{i},j\in(S_{\beta})_{i^{\prime}}}(x^{i}_{l}-x^{i^{\prime}}_{j})^{\kappa_{ii^{\prime}}}\prod_{i\in I,l\neq j,l,j\in(S_{\beta})_{i}}(x^{i}_{l}-x^{i}_{j})^{\kappa_{ii}},

where for an arrow aEa\in E, s(a)s(a) is the source and t(a)t(a) is the target. The difference between l(β)l(\beta) and lQ(β)l_{Q}(\beta) is possibly a sign which depends on the directions of the arrows in QQ. So accordingly, in the formula (13) of sSs^{S}, we replace l(S)l(S) by

lQ(S)=aE such that s(a)=i,t(a)=i,ii,lSi,jSi(xlixji)κiiiI,lj,l,jSi(xlixji)κii.l_{Q}(S)=\prod_{a\in E\text{ such that }s(a)=i,t(a)=i^{\prime},i\neq i^{\prime},l\in S_{i},j\in S_{i^{\prime}}}(x^{i}_{l}-x^{i^{\prime}}_{j})^{\kappa_{ii^{\prime}}}\prod_{i\in I,l\neq j,l,j\in S_{i}}(x^{i}_{l}-x^{i}_{j})^{\kappa_{ii}}.

The relations in the algebra A0,regA_{0,reg} for sS,SSαs^{S},S\subset S^{\alpha} are

sAsB=lQ(A)lQ(B)lQ(AB)lQ(AB)sABsAB.s^{A}s^{B}=\frac{l_{Q}(A)l_{Q}(B)}{l_{Q}(A\cup B)l_{Q}(A\cap B)}s^{A\cup B}s^{A\cap B}.

Now to check sS,SSαs^{S},S\subset S^{\alpha} and xχ(S)t,SSαx^{\chi(S)}t,S\subset S^{\alpha} satisfy the same relations, it suffices to show that

fcQ(A,B)fcQ(AB,AB)Eu(AB)Eu(AB)Eu(A)Eu(B)=lQ(A)lQ(B)lQ(AB)lQ(AB).\frac{\text{fc}_{Q}(A,B)}{\text{fc}_{Q}(A\cup B,A\cap B)}\frac{Eu(A\cup B)Eu(A\cap B)}{Eu(A)Eu(B)}=\frac{l_{Q}(A)l_{Q}(B)}{l_{Q}(A\cup B)l_{Q}(A\cap B)}. (24)

This will be an elementary combinatorial calculation. For a general quiver Q,Q, denote by QsQ_{s} the quiver removing all arrows between different vertices and QbQ_{b} the removing all arrows.

  • We first prove the case where QQ has no edges, i.e. Q=Qb.Q=Q_{b}. In this case N=0N=0, so

    fcQb(A,B)fcQb(AB,AB)=1.\frac{\text{fc}_{Q_{b}}(A,B)}{\text{fc}_{Q_{b}}(A\cup B,A\cap B)}=1.

    It suffices to show that

    Eu(AB)Eu(AB)Eu(A)Eu(B)=lQb(A)lQb(B)lQb(AB)lQb(AB).\frac{Eu(A\cup B)Eu(A\cap B)}{Eu(A)Eu(B)}=\frac{l_{Q_{b}}(A)l_{Q_{b}}(B)}{l_{Q_{b}}(A\cup B)l_{Q_{b}}(A\cap B)}. (25)

    From the description of Eu(S)Eu(S) and lQb(S)l_{Q_{b}}(S), we can assume II is one point. Let W:={1,2,,n}W:=\{1,2,\cdots,n\}. Let AB=C,AB=D,BA=E,W(AB)=FA\cap B=C,A\setminus B=D,B\setminus A=E,W\setminus(A\cup B)=F so A=CD,B=DE,AB=CDE,AB=D,W=CDEF.A=C\cup D,B=D\cup E,A\cup B=C\cup D\cup E,A\cap B=D,W=C\cup D\cup E\cup F. By lemma 12(b), we have

    Eu1(A)=lA,jWA(alaj)=(l,j)A×(WA)(alaj).Eu^{-1}(A)=\prod_{l\in A,j\in W\setminus A}(a_{l}-a_{j})=\prod_{(l,j)\in A\times(W\setminus A)}(a_{l}-a_{j}).

    Here, we rewrite the index set in the second equality as a product A×(WA)A\times(W\setminus A). It is clear to show 25, it suffices to compare the index set of the products. We write

    X=suppSX\overset{\text{supp}}{=}S

    to mean

    X=(l,j)S(alaj).X=\prod_{(l,j)\in S}(a_{l}-a_{j}).

    By this notation,

    Eu1(A)=suppA×(WA)=(CD)×(EF)=(C×E)(C×F)(D×E)(D×F).Eu^{-1}(A)\overset{\text{supp}}{=}A\times(W\setminus A)=(C\cup D)\times(E\cup F)=(C\times E)\cup(C\times F)\cup(D\times E)\cup(D\times F).
    Eu1(B)=lB,jWB(alaj)=supp(D×C)(D×F)(E×C)(E×F).Eu^{-1}(B)=\prod_{l\in B,j\in W\setminus B}(a_{l}-a_{j})\overset{\text{supp}}{=}(D\times C)\cup(D\times F)\cup(E\times C)\cup(E\times F).
    Eu1(AB)=lAB,jW(AB)(alaj)=supp(C×F)(D×F)(E×F)Eu^{-1}(A\cup B)=\prod_{l\in A\cup B,j\in W\setminus(A\cup B)}(a_{l}-a_{j})\overset{\text{supp}}{=}(C\times F)\cup(D\times F)\cup(E\times F)

    Eu1(AB)=lAB,jW(AB)(alaj)=supp(D×C)(D×E)(D×F).Eu^{-1}(A\cap B)=\prod_{l\in A\cap B,j\in W\setminus(A\cap B)}(a_{l}-a_{j})\overset{\text{supp}}{=}(D\times C)\cup(D\times E)\cup(D\times F).

    So

    Eu1(A)Eu1(B)Eu1(AB)Eu1(aB)=supp(C×E)(E×C).\frac{Eu^{-1}(A)Eu^{-1}(B)}{Eu^{-1}(A\cup B)Eu^{-1}(a\cap B)}\overset{\text{supp}}{=}(C\times E)\cup(E\times C).

    For lQbl_{Q_{b}}, we have

    lQb(A)=lA,jA(alaj)=supp{(C×C)(C×D)(D×C)(D×D)}.l_{Q_{b}}(A)=\prod_{l\in A,j\in A}(a_{l}-a_{j})\overset{\text{supp}}{=}\{(C\times C)\cup(C\times D)\cup(D\times C)\cup(D\times D)\}.
    lQb(B)=lB,jB(alaj)=supp{(D×D)(D×E)(E×D)(E×E)}.l_{Q_{b}}(B)=\prod_{l\in B,j\in B}(a_{l}-a_{j})\overset{\text{supp}}{=}\{(D\times D)\cup(D\times E)\cup(E\times D)\cup(E\times E)\}.
    lQb(AB)=lAB,jAB(alaj)=suppl_{Q_{b}}(A\cup B)=\prod_{l\in A\cup B,j\in A\cup B}(a_{l}-a_{j})\overset{\text{supp}}{=}\\
    {(C×C)(C×D)(C×E)(D×C)(D×D)(D×E)(E×C)(E×D)(E×E)}.\{(C\times C)\cup(C\times D)\cup(C\times E)\cup(D\times C)\cup(D\times D)\cup(D\times E)\cup(E\times C)\cup(E\times D)\cup(E\times E)\}.
    lQb(AB)=lAB,jAB(alaj)=supp{(D×D)}.l_{Q_{b}}(A\cap B)=\prod_{l\in A\cap B,j\in A\cap B}(a_{l}-a_{j})\overset{\text{supp}}{=}\{(D\times D)\}.

    So

    lQb(A)lQb(B)lQb(AB)lQb(AB)=supp(C×E)(E×C)\frac{l_{Q_{b}}(A)l_{Q_{b}}(B)}{l_{Q_{b}}(A\cup B)l_{Q_{b}}(A\cap B)}\overset{\text{supp}}{=}(C\times E)\cup(E\times C)

    and therefore (25) holds.

  • Now we prove the case where Q=QsQ=Q_{s} has only self-edges. Again it is clear we can assume II is one point. Denote

    lQsd=lQslQb.l^{d}_{Q_{s}}=\frac{l_{Q_{s}}}{l_{Q_{b}}}.

    Now it suffices to show that

    fcQs(A,B)fcQs(AB,AB)=lQsd(A)lQsd(B)lQsd(AB)lQsd(AB).\frac{\text{fc}_{Q_{s}}(A,B)}{\text{fc}_{Q_{s}}(A\cup B,A\cap B)}=\frac{l^{d}_{Q_{s}}(A)l^{d}_{Q_{s}}(B)}{l^{d}_{Q_{s}}(A\cup B)l^{d}_{Q_{s}}(A\cap B)}.

    By remark 4, we can reduce to the case where QsQ_{s} has one self-loop and it suffices to prove

    fcQs(A,B)fcQs(AB,AB)=supp(C×E)(E×C).\frac{\text{fc}_{Q_{s}}(A,B)}{\text{fc}_{Q_{s}}(A\cup B,A\cap B)}\overset{\text{supp}}{=}(C\times E)\cup(E\times C).

    Now N=Hom(V,V)N=Hom(V,V) is the adjoint representation of GL(V).GL(V). The characters ξi\xi_{i} is the set of all roots of GL(V),GL(V), which are all (alaj)(a_{l}-a_{j}) for lj,l,jW.l\neq j,l,j\in W. Compute the pairing between A,B,AB,ABA,B,A\cup B,A\cap B and (alaj)(a_{l}-a_{j}) for lj,l,jWl\neq j,l,j\in W and A,BA,B we get the formula.

  • Now we prove the general case. Denote lQd=lQlQsl^{d}_{Q}=\frac{l_{Q}}{l_{Q_{s}}} and fcQd=fcQfcQs.\text{fc}^{d}_{Q}=\frac{\text{fc}_{Q}}{\text{fc}_{Q_{s}}}. It suffices to show that

    fcQd(A,B)fcQd(AB,AB)=lQd(A)lQd(B)lQd(AB)lQd(AB).\frac{\text{fc}^{d}_{Q}(A,B)}{\text{fc}^{d}_{Q}(A\cup B,A\cap B)}=\frac{l^{d}_{Q}(A)l^{d}_{Q}(B)}{l^{d}_{Q}(A\cup B)l^{d}_{Q}(A\cap B)}.

    Again by remark 4, we can reduce to the case where QQ has two vertices and only one arrow 12.1\xrightarrow[]{}2. In this case N=Hom(V1,V2)N=Hom(V_{1},V_{2}). The characters ξi\xi_{i} consist of (al1aj2)(a^{1}_{l}-a^{2}_{j}) for all lln1,1jn2.l\leq l\leq n_{1},1\leq j\leq n_{2}. Compute the pairing between A,B,AB,ABA,B,A\cup B,A\cap B and (al1aj2)(a^{1}_{l}-a^{2}_{j}) for all lln1,1jn2l\leq l\leq n_{1},1\leq j\leq n_{2} and A,BA,B we get the formula.

5 Appendix: Proof of lemma 4 (by Ivan Mirković)

5.1 Equations of Segre embeddings

The locality equations in the discrete range are just the Segre embedding equations of a special type (1)n2n1(\mathbb{P}^{1})^{n}\hookrightarrow\mathbb{P}^{2^{n}-1}. The standard list of Segre equations of general type pD(Vp)(VD)\prod_{p\in D}\mathbb{P}(V_{p})\hookrightarrow\mathbb{P}(V_{D}), is checked based on the case (A)×(B)(AB)\mathbb{P}(A)\times\mathbb{P}(B)\hookrightarrow\mathbb{P}(A\otimes B) in 5.1.1. In our case (1)n2n1(\mathbb{P}^{1})^{n}\hookrightarrow\mathbb{P}^{2^{n}-1} the combinatorics is stated in terms of Gr(D)[D]Gr(D)\subset\mathbb{N}[D] in 5.1.2.

5.1.1 Segre embeddings.

(i) Case of two factors. A choice of coordinates on AA and BB, xa,a𝒜x_{a},a\in\mathcal{A} and yb,b,y_{b},b\in\mathcal{B}, gives coordinates zab=xaybz_{ab}=x_{a}\otimes y_{b} on ABA\otimes B. A vector vABv\in A\otimes B can be thought of as an operator v:ABv:A^{*}\xrightarrow[]{}B and vv is a pure tensor iff the rank of vv is 1.\leq 1. In terms of the matrix (zab)𝒜×(z_{ab})_{\mathcal{A}\times\mathcal{B}} this condition is the vanishing of all of its 2×22\times 2 minors zabzabzabzab.z_{ab}z_{a^{\prime}b^{\prime}}-z_{ab^{\prime}}z_{a^{\prime}b}.
(ii) Segre embedding map pD(Vp)(VD).\prod_{p\in D}\mathbb{P}(V_{p})\xrightarrow[]{}\mathbb{P}(V_{D}). Let xip,ip,x^{p}_{i},i\in\mathcal{B}_{p}, be the coordinates on vector spaces Vp,pD,V_{p},p\in D, so that for XDX\subset D on VX=defpXVp.V_{X}\stackrel{{\scriptstyle\text{def}}}{{=}}\otimes_{p\in X}V_{p}. We have coordinates zβX=defpDxβppz_{\beta}^{X}\stackrel{{\scriptstyle\text{def}}}{{=}}\otimes_{p\in D}x_{\beta_{p}}^{p} indexed by βX=defpXp.\beta\in\mathcal{B}_{X}\stackrel{{\scriptstyle\text{def}}}{{=}}\prod_{p\in X}\mathcal{B}_{p}. Then the Segre embedding map pD(Vp)(VD)\prod_{p\in D}\mathbb{P}(V_{p})\hookrightarrow\mathbb{P}(V_{D}) is given by zβDpDxβpp.z_{\beta}^{D}\mapsto\prod_{p\in D}x_{\beta_{p}}^{p}. We view pD(Vp)\prod_{p\in D}\mathbb{P}(V_{p}) as the intersection of all (VX)×(VY)(VD)\mathbb{P}(V_{X})\times\mathbb{P}(V_{Y})\subset\mathbb{P}(V_{D}) over all decomposition D=XY.D=X\sqcup Y.232323The claim (A)×(BC)(A×B×C)(A×B)×(C)=(A)×(B)×(C)\mathbb{P}(A)\times\mathbb{P}(B\otimes C)\cap_{\mathbb{P}(A\times B\times C)}\mathbb{P}(A\times B)\times\mathbb{P}(C)=\mathbb{P}(A)\times\mathbb{P}(B)\times\mathbb{P}(C) means that if a vector vv in V=ABCV=A\otimes B\otimes C is a pure tensor for A(BC)A\otimes(B\otimes C) and (AB)C(A\otimes B)\otimes C, then it is also a pure tensor for ABC.A\otimes B\otimes C. Proof. The assumption is v=aα=γcv=a\otimes\alpha=\gamma\otimes c with aA,cC,αBC,γAB.a\in A,c\in C,\alpha\in B\otimes C,\gamma\in A\otimes B. Choose a basis cic_{i} of CC with a dual basis cic^{i} and c=c0c=c_{0}. For i0i\neq 0 we have cic,c^{i}\perp c, hence 0=v,ci=abi0=\langle v,c^{i}\rangle=a\otimes b_{i} hence bi=0b_{i}=0. So, v=ab0c.v=a\otimes b_{0}\otimes c. So, by (i) equations are the minors |zαβzαβ′′zα′′βzα′′β′′|\begin{vmatrix}z_{\alpha^{\prime}\beta^{\prime}}&z_{\alpha^{\prime}\beta^{\prime\prime}}\\ z_{\alpha^{\prime\prime}\beta^{\prime}}&z_{\alpha^{\prime\prime}\beta^{\prime\prime}}\end{vmatrix} indexed by data D=XYD=X\cup Y and α,α′′X,β,β′′Y\alpha^{\prime},\alpha^{\prime\prime}\in\mathcal{B}_{X},\beta^{\prime},\beta^{\prime\prime}\in\mathcal{B}_{Y} (so that αβD\alpha^{\prime}\beta^{\prime}\in\mathcal{B}_{D} etc.).

Remark 7.

We can write these equations as zϕzψzϕCzψCz_{\phi}z_{\psi}-z_{\phi^{C}}z_{\psi^{C}} for (ϕ,ψ,C)D2×Gr(D);(\phi,\psi,C)\in\mathcal{B}_{D}^{2}\times Gr(D); here CGr(D)C\in Gr(D) acts on D2\mathcal{B}_{D}^{2} by the unique involution σC(ϕ,ψ)=(ϕC,ψC)\sigma_{C}(\phi,\psi)=(\phi^{C},\psi^{C}) that exchanges the values on CC, i.e., ϕpC=ψp\phi^{C}_{p}=\psi_{p} for pCp\in C and ψpC=ψp\psi^{C}_{p}=\psi_{p} for pCp\notin C (and the same for ψC\psi^{C}). So, the equation require invariance of products under Gr(D)Gr(D) as a DD-degeneration of commutativity which is the case C=DC=D as (ϕD,ψD)=(ϕ,ψ).(\phi^{D},\psi^{D})=(\phi,\psi). (However, this Gr(D)Gr(D) action is only defined on a chosen basis.)

5.1.2 The case of (1)n2n1.(\mathbb{P}^{1})^{n}\hookrightarrow\mathbb{P}^{2^{n}-1}.

Now we have identifications p{,p}=Gr(p)\mathcal{B}_{p}\xleftarrow[\cong]{}\{\emptyset,p\}=Gr(p) and therefore DGr(D)\mathcal{B}_{D}\xleftarrow[\cong]{}Gr(D) (by ϕX\phi\mapsfrom X for ϕp=δpX).\phi_{p}=\delta_{p\in X}). We also embed Gr(D)Gr(D) into a monoid [D].\mathbb{N}[D]. (We often denote zXz_{X} by 1X1_{X}.)

Corollary 4.

The Segre equations are now zXzY=zUzVz_{X}z_{Y}=z_{U}z_{V}, indexed by all X,Y,U,VDX,Y,U,V\subset D with X+Y=U+V[D].X+Y=U+V\in\mathbb{N}[D].

Proof.

Involution σC,CD\sigma_{C},C\subset D preserve for each pDp\in D the multiset ϕp,ψp\phi_{p},\psi_{p} hence also the sum ϕp+ψp.\phi_{p}+\psi_{p}. Conversely, if X+Y=U+VX+Y=U+V then (U,V)=σC(X,Y)(U,V)=\sigma_{C}(X,Y) for C={pD;XpUp}.C=\{p\in D;X_{p}\neq U_{p}\}.

Remark 8.

(0) One has zi,jz=zizjz_{i,j}z_{\emptyset}=z_{i}z_{j} and inductively zXz|X|1=pDzp.z_{X}z_{\emptyset}^{|X|-1}=\prod_{p\in D}z_{p}. (1) By setting 1=11_{\emptyset}=1 one obtains an open affine subspace of (VD)\mathbb{P}(V_{D}) with functions 𝕜[1B,BD]\mathbb{k}[1_{B},\emptyset\neq B\subset D]. Here, locality equations reduce to zX=pDzpz_{X}=\prod_{p\in D}z_{p} with solutions 𝔸D.\mathbb{A}^{D}.

References

  • [AR] Pramod N. Achar and Simon Riche. Central sheaves on affine flag varieties.
  • [BF10] Alexander Braverman and Michael Finkelberg. Pursuing the double affine grassmannian I: Transversal slices via instantons on AkA_{k}-singularities. Duke Math. J., 152(arXiv: 0711.2083):175–206, 2010.
  • [BF14] Alexander Braverman and Michael Finkelberg. Semi-infinite schubert varieties and quantum K-theory of flag manifolds. Journal of the American Mathematical Society, 27(4):1147–1168, 2014.
  • [BFG06] Alexander Braverman, Michael Finkelberg, and Dennis Gaitsgory. Uhlenbeck spaces via affine lie algebras. In The unity of mathematics: In Honor of the Ninetieth Birthday of IM Gelfand, pages 17–135. Springer, 2006.
  • [BFN18] Alexander Braverman, Michael Finkelberg, and Hiraku Nakajima. Towards a mathematical definition of coulomb branches of 33-dimensional n=4n=4 gauge theories, II. Advances in Theoretical and Mathematical Physics, 22(5):1071–1147, 2018.
  • [BFN19] Alexander Braverman, Michael Finkelberg, and Hiraku Nakajima. Coulomb branches of 3d3d n=4n=4 quiver gauge theories and slices in the affine Grassmannian. ADV. THEOR. MATH. PHYS, 23(1):75–166, 2019.
  • [Don] Zhijie Dong. On the relations of the generators of the homogenous coordinate ring of zastava for matrices.
  • [FKM20] Michael Finkelberg, Vasily Krylov, and Ivan Mirković. Drinfeld–gaitsgory–vinberg interpolation grassmannian and geometric satake equivalence. Journal of Topology, 13(2):683–729, 2020.
  • [FM99] Michael Finkelberg and Ivan Mirković. Semi-infinite flags I. Case of global curve 1\mathbb{P}^{1}. Differential Topology, Infinite-Dimensional Lie Algebras, and Applications, pages 81–112, 1999.
  • [Mir23] Ivan Mirković. Lattice vertex algebras and Loop Grassmannians. Transformation Groups, 28(3):1221–1243, 2023.
  • [MW24] Dinakar Muthiah and Alex Weekes. Fundamental monopole operators and embeddings of Kac-Moody affine Grassmannian slices. International Mathematics Research Notices, page rnae115, 2024.
  • [MYZ21] Ivan Mirković, Yaping Yang, and Gufang Zhao. Loop Grassmannians of quivers and affine quantum groups. In Representation Theory and Algebraic Geometry: A Conference Celebrating the Birthdays of Sasha Beilinson and Victor Ginzburg, pages 347–392. Springer, 2021.
  • [Ras21] Sam Raskin. Chiral principal series categories i: finite dimensional calculations. Advances in Mathematics, 388:107856, 2021.
  • [Zhu09] Xinwen Zhu. Affine demazure modules and t-fixed point subschemes in the affine grassmannian. Advances in Mathematics, 221(2):570–600, 2009.