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

\autonum@generatePatchedReferenceCSL

Cref

Algorithmic aspects of semistability of quiver representations

Yuni Iwamasa Graduate School of Informatics, Kyoto University, Kyoto, 606-8501, Japan. Email: iwamasa@i.kyoto-u.ac.jp    Taihei Oki Institute for Chemical Reaction Design and Discovery (ICReDD), Hokkaido University, Sapporo, 001-0021, Japan. Email: oki@icredd.hokudai.ac.jp    Tasuku Soma The Institute of Statistical Mathematics, Tokyo, 190-8562, Japan. Email: soma@ism.ac.jp
Abstract

We study the semistability of quiver representations from an algorithmic perspective. We present efficient algorithms for several fundamental computational problems on the semistability of quiver representations: deciding the semistability and σ\sigma-semistability, finding the maximizers of King’s criterion, and computing the Harder–Narasimhan filtration. We also investigate a class of polyhedral cones defined by the linear system in King’s criterion, which we refer to as King cones. For rank-one representations, we demonstrate that these King cones can be encoded by submodular flow polytopes, enabling us to decide the σ\sigma-semistability in strongly polynomial time. Our approach employs submodularity in quiver representations, which may be of independent interest.

1 Introduction and our contribution

Quiver representation is a simple generalization of matrices that has led to surprisingly deep extensions of various results in linear algebra [DW17]. In this paper, we study the semistability of quiver representations, which is a central concept in the geometric invariant theory (GIT), from an algorithmic perspective. The semistability of quiver representations appears in operator scaling [Gur04, GGOW19, Fra18, BFGO+18, FSG23], Brascamp–Lieb (BL) inequality [GGOW18], Tyler’s M-estimator [FM20], and scatter estimation of structured normal models [AKRS21], which have attracted considerable attention in theoretical computer science owing to their connection to the noncommutative Edmonds’ problem, algebraic complexity theory, and submodular optimization [Mul17, IQS18, BFGO+19, HH21]. The goal of this paper is to provide efficient algorithms for various fundamental computational problems on the semistability of quiver representations. In the following, we describe the problems more formally and present our results.

1.1 Semistability of quiver representations

Here, we present the formal definition of a quiver representation. We follow the terminologies in [DW17, BFGO+19]. Let Q=(Q0,Q1)Q=(Q_{0},Q_{1}) be a quiver with a vertex set Q0Q_{0} and an arc set Q1Q_{1}. In this paper, we consider only acyclic quivers except in Section 6. For each arc aQ1a\in Q_{1}, we denote the tail and head of aa by tata and haha, respectively. A representation VV of QQ consists of complex vector spaces V(i)V(i) for vertex iQ0i\in Q_{0} and linear maps V(a):V(ta)V(ha)V(a):V(ta)\to V(ha) for arc aQ1a\in Q_{1}. A subrepresentation WW of VV is a representation of the same quiver such that W(i)V(i)W(i)\leq V(i) for iQ0i\in Q_{0}, and W(a)=V(a)|W(ta)W(a)=V(a)|_{W(ta)} and imW(a)W(ha)\operatorname{im}W(a)\leq W(ha) for aQ1a\in Q_{1}. The vector of dimensions of V(i)V(i) is called the dimension vector of the representation, denoted by dim¯V\operatorname{\underline{dim}}V. We call the vector space of all representations of QQ with the dimension vector α\alpha the representation space of QQ with dimension vector α\alpha, which we denote by Rep(Q,α)\operatorname{Rep}(Q,\alpha). After fixing the dimension vector α\alpha and a basis of each V(i)V(i), we can represent V(a)V(a) as an α(ha)×α(ta)\alpha(ha)\times\alpha(ta) matrix. Therefore, the representation space can be identified as

Rep(Q,α)=aQ1Mat(α(ha),α(ta)),\displaystyle\operatorname{Rep}(Q,\alpha)=\bigoplus_{a\in Q_{1}}\operatorname{Mat}(\alpha(ha),\alpha(ta)), (1.1)

where Mat(m,n)\operatorname{Mat}(m,n) denotes the space of m×nm\times n complex matrices.

Fix a dimension vector α\alpha. Let

GL(Q,α)iQ0GL(α(i)),\displaystyle\operatorname{GL}(Q,\alpha)\coloneqq\prod_{i\in Q_{0}}\operatorname{GL}(\alpha(i)), (1.2)

where GL(n)\operatorname{GL}(n) denotes the general linear group of degree nn. Then, GL(Q,α)\operatorname{GL}(Q,\alpha) acts on the representation space by a change of basis:

gV(ghaV(a)gta1)aQ1.\displaystyle g\cdot V\coloneqq(g_{ha}V(a)g_{ta}^{-1})_{a\in Q_{1}}. (1.3)

Note that this is a left action, i.e., (gh)V=g(hV)(gh)\cdot V=g\cdot(h\cdot V) for g,hGL(Q,α)g,h\in\operatorname{GL}(Q,\alpha). We say that a representation VV is semistable under the GL(Q,α)\operatorname{GL}(Q,\alpha)-action if the orbit closure of VV does not contain the origin, i.e.,

infgGL(Q,α)aQ1ghaV(a)gta1F2>0.\displaystyle\inf_{g\in\operatorname{GL}(Q,\alpha)}\sum_{a\in Q_{1}}\lVert g_{ha}V(a)g_{ta}^{-1}\rVert_{\mathrm{F}}^{2}>0. (1.4)

Otherwise, VV is said to be unstable. The set of all unstable representations is called the null-cone of the GL(Q,α)\operatorname{GL}(Q,\alpha)-action. It is easy to see that any representation is unstable under the GL(Q,α)\operatorname{GL}(Q,\alpha)-action if QQ is acyclic.111The readers may wonder whether we can check the semistability (under GL(Q,α)\operatorname{GL}(Q,\alpha)-action) of quiver representations of cyclic quivers. We will address this point later.

However, the semistability of quivers under subgroups of GL(Q,α)\operatorname{GL}(Q,\alpha) turns out to be more intricate. Let σQ0\sigma\in\mathbb{Z}^{Q_{0}} be an integer vector on Q0Q_{0}, which we call a weight. Let χσ\chi_{\sigma} be the corresponding multiplicative character of GL(Q,α)\operatorname{GL}(Q,\alpha), i.e.,

χσ(g)=iQ0det(gi)σ(i).\chi_{\sigma}(g)=\prod_{i\in Q_{0}}\det(g_{i})^{\sigma(i)}.

Note that χσ\chi_{\sigma} is a one-dimensional representation of GL(Q,α)\operatorname{GL}(Q,\alpha); GL(Q,α)\operatorname{GL}(Q,\alpha) acts on \mathbb{C} by gxχσ(g)xg\cdot x\coloneqq\chi_{\sigma}(g)x. A representation VV is said to be σ\sigma-semistable if the orbit closure of (V,1)Rep(Q,α)(V,1)\in\operatorname{Rep}(Q,\alpha)\oplus\mathbb{C} under the GL(Q,α)\operatorname{GL}(Q,\alpha) action does not contain the origin, i.e.,

infgGL(Q,α)(aQ1ghaV(a)gta1F2+|χσ(g)|2)>0.\displaystyle\inf_{g\in\operatorname{GL}(Q,\alpha)}\left(\sum_{a\in Q_{1}}\lVert g_{ha}V(a)g_{ta}^{-1}\rVert_{\mathrm{F}}^{2}+\lvert\chi_{\sigma}(g)\rvert^{2}\right)>0. (1.5)

It turns out that checking the σ\sigma-semistability of a quiver representation includes operator scaling (noncommutative rank computation) and the membership problem of the BL polytopes. We will see these examples in the following sections.

Our first result is a deterministic algorithm that, given a quiver representation VV and weight σ\sigma, decides whether the representation is σ\sigma-semistable in time polynomial in the bit complexity of VV and absolute values of the entries of σ\sigma. Let α(Q0)iQ0α(i)\alpha(Q_{0})\coloneqq\sum_{i\in Q_{0}}\alpha(i).

Theorem 1.1 (informal version of Theorem 3.4).

Let QQ be an acyclic quiver, VV a representation of QQ, and σ\sigma a weight. There is a deterministic algorithm that decides the σ\sigma-semistability of VV in time polynomial in the size of QQ, α(Q0)\alpha(Q_{0}), bit complexity of VV, and absolute values of the entries of σ\sigma.

This improves the previous result [Hus21] which runs in time polynomial in the number of paths in QQ, which can be exponential in the size of QQ. Furthermore, if the absolute value of the entries of σ\sigma is constant, our algorithm runs in polynomial time. This includes the known result for operator scaling [GGOW19].

1.2 King’s criterion

[Kin94] showed the following characterization of σ\sigma-semistability, which is known as King’s criterion. Let σ(α)iQ0σ(i)α(i)\sigma(\alpha)\coloneqq\sum_{i\in Q_{0}}\sigma(i)\alpha(i) for a dimension vector α\alpha. Then, a representation VV is σ\sigma-semistable if and only if σ(dim¯V)=0\sigma(\operatorname{\underline{dim}}V)=0 and σ(dim¯W)0\sigma(\operatorname{\underline{dim}}W)\leq 0 for any subrepresentation WW of VV.

King’s criterion is a common generalization of the noncommutative rank (nc-rank) computation and membership problem of BL polytopes.

Example 1.2 (nc-rank).

Let QQ be the generalized Kronecker quiver with mm parallel arcs, α=(n,n)\alpha=(n,n), and σ=(1,1)\sigma=(1,-1); see Figure 1. Any representation VV of QQ with the dimension vector α\alpha can be regarded as an n×nn\times n linear matrix A=a=1mxaV(a)A=\sum_{a=1}^{m}x_{a}V(a), where xax_{a} is an indeterminate. A subrepresentation WW of VV is determined by a pair of subspaces (W(1),W(2))(W(1),W(2)) such that a=1mV(a)W(1)W(2)\sum_{a=1}^{m}V(a)W(1)\leq W(2). Then, King’s criterion reads that VV is σ\sigma-semistable if and only if dimUdim(a=1mV(a)U)0\dim U-\dim(\sum_{a=1}^{m}V(a)U)\leq 0 for any subspace UnU\leq\mathbb{C}^{n}, which is equivalent to that AA is nc-nonsingular. More generally, the nc-rank of AA is equal to the minimum of n+dimUdim(a=1mV(a)U)n+\dim U-\dim(\sum_{a=1}^{m}V(a)U) over all subspaces UnU\leq\mathbb{C}^{n} [FR04].

Example 1.3 (BL polytope).

Let QQ be a star quiver with mm leaves. We assume that Q0={0,1,,m}Q_{0}=\{0,1,\dots,m\} and 0 is the root; see Figure 1. Let α=(n,n1,,nm)\alpha=(n,n_{1},\dots,n_{m}) and σ=(d,c1,,cm)\sigma=(d,-c_{1},\dots,-c_{m}) for positive integers d,c1,,cmd,c_{1},\dots,c_{m}. A real representation VV of QQ with the dimension vector α\alpha can be regarded as a tuple of the matrices (B1,,Bm)(B_{1},\dots,B_{m}), where BiB_{i} is an ni×nn_{i}\times n matrix. Again, a subrepresentation WW is an (m+1)(m+1)-tuple of the subspaces (W(0),W(1),,W(m))(W(0),W(1),\dots,W(m)) such that BiW(0)W(i)B_{i}W(0)\leq W(i) for i[m]i\in[m]. King’s criterion reads that VV is σ\sigma-semistable if and only if dn=i=1mcinidn=\sum_{i=1}^{m}c_{i}n_{i} and dndimW(0)i=1mcinidim(BiW(0))0dn\dim W(0)-\sum_{i=1}^{m}c_{i}n_{i}\dim(B_{i}W(0))\leq 0 for any subspace W(0)W(0). This is equivalent to that p=(c1/d,,cm/d)p=(c_{1}/d,\dots,c_{m}/d) is in the BL polytope of linear operators B1,,BmB_{1},\dots,B_{m} [BCCT08].

Refer to caption
Figure 1: Generalized Kronecker quiver (left) and star quiver (right).

We study the following optimization problem: given a quiver representation VV and weight σ\sigma, find a subrepresentation WW of VV that maximizes σ(dim¯W)\sigma(\operatorname{\underline{dim}}W). In the case of the nc-rank, such a subrepresentation corresponds to a subspace UU that maximizes dimUdim(a=1mAiU)\dim U-\dim(\sum_{a=1}^{m}A_{i}U). Such a subspace is called a shrunk subspace and can be regarded as a certificate of the nc-rank [IQS18, FSG23]. In the case of the BL polytopes, the problem corresponds to separation for the BL polytope [GGOW18].

King’s criterion can be regarded as maximizing a modular function over the modular lattice of subrepresentations. For any subrepresentations W1,W2W_{1},W_{2} of VV, define the subrepresentations W1+W2W_{1}+W_{2} and W1W2W_{1}\cap W_{2} as follows. For each iQ0i\in Q_{0},

(W1+W2)(i)W1(i)+W2(i),(W1W2)(i)W1(i)W2(i),\displaystyle(W_{1}+W_{2})(i)\coloneqq W_{1}(i)+W_{2}(i),\quad(W_{1}\cap W_{2})(i)\coloneqq W_{1}(i)\cap W_{2}(i), (1.6)

where the addition and intersection on the right-hand side are those of the vector spaces. Furthermore, the linear map of aQ1a\in Q_{1} in W1+W2W_{1}+W_{2} (resp. W1W2W_{1}\cap W_{2}) is defined as the restriction of V(a)V(a) to (W1+W2)(ta)(W_{1}+W_{2})(ta) (resp. (W1W2)(ta)(W_{1}\cap W_{2})(ta)). Then, W1+W2W_{1}+W_{2} and W1W2W_{1}\cap W_{2} are indeed subrepresentations of VV. Thus, the subrepresentations of VV form a modular lattice. Furthermore, the function f(W)σ(dim¯W)f(W)\coloneqq\sigma(\operatorname{\underline{dim}}W) is a modular function, i.e., for any subrepresentations W1,W2W_{1},W_{2} of VV,

f(W1)+f(W2)=f(W1+W2)+f(W1W2).\displaystyle f(W_{1})+f(W_{2})=f(W_{1}+W_{2})+f(W_{1}\cap W_{2}). (1.7)

Thus, subrepresentations maximizing ff form a sublattice, and there is a unique inclusion-wise minimum maximizer. Our second result is a deterministic algorithm to find such a maximizer of King’s criterion.

Theorem 1.4 (informal version of Theorem 3.5).

Let QQ be an acyclic quiver, VV a representation of QQ, and σ\sigma a weight. There is a deterministic algorithm that finds the inclusion-wise minimum maximizer WW of King’s criterion in time polynomial in the size of QQ, α(Q0)\alpha(Q_{0}), bit complexity of VV, and absolute values of the entries of σ\sigma.

King’s criterion was originally proved using the Hilbert-Mumford criterion (see, e.g., [DW17, Sections 9.6 and 9.8]), a fundamental result in the GIT. We provide an alternative elementary proof in Appendix A for the sake of completeness.

1.3 Harder-Narasimhan filtration

We use the algorithm for finding the maximizers of King’s criterion to devise an algorithm for finding the Harder-Narasimhan (HN) filtration [HN75, HA02] of a quiver representation. Roughly speaking, the HN-filtration decomposes a quiver representation into the direct sum of smaller representations.

More precisely, let σQ0\sigma\in\mathbb{Z}^{Q_{0}} be a weight and τ+Q0\tau\in\mathbb{Z}_{+}^{Q_{0}} a strictly monotone weight, i.e., a nonnegative weight such that τ(dim¯W)>0\tau(\operatorname{\underline{dim}}W)>0 if W{0}W\neq\{0\}. Here, {0}\{0\} denotes the zero representation, which is the representation whose dimension vector is the zero vector. We define the slope of a quiver nonzero representation VV as μ(V)=σ(dim¯V)/τ(dim¯V)\mu(V)=\sigma(\operatorname{\underline{dim}}V)/\tau(\operatorname{\underline{dim}}V). We say that VV is μ\mu-semistable222It is also called (σ:τ)(\sigma:\tau)-semistability in the literature. if μ(W)μ(V)\mu(W)\leq\mu(V) for any nonzero subrepresentation WW of VV. The HN-filtration theorem states that for any quiver representation VV, there exists a unique filtration {0}=W0<W1<<Wk=V\{0\}=W_{0}<W_{1}<\cdots<W_{k}=V such that (i) μ(Wi/Wi1)>μ(Wi+1/Wi)\mu(W_{i}/W_{i-1})>\mu(W_{i+1}/W_{i}) for i[k1]i\in[k-1] and (ii) Wi/Wi1W_{i}/W_{i-1} is μ\mu-semistable. Here, Wi<Wi+1W_{i}<W_{i+1} means that WiW_{i} is a subrepresentation of Wi+1W_{i+1} with WiWi+1W_{i}\neq W_{i+1}, and Wi/Wi1W_{i}/W_{i-1} is a representation of QQ such that (Wi/Wi1)(j)(W_{i}/W_{i-1})(j) is the quotient space Wi(j)/Wi1(j)W_{i}(j)/W_{i-1}(j) for jQ0j\in Q_{0} and (Wi/Wi1)(a)(W_{i}/W_{i-1})(a) is the corresponding quotient linear map of Wi(a)W_{i}(a) for aQ1a\in Q_{1}. We note that semistability with respect to a slope can be reduced to that for a weight; see Lemma 4.4.

Our third result is a deterministic algorithm for finding the HN-filtration.

Theorem 1.5 (informal version of Theorem 4.5).

Let QQ be an acyclic quiver, VV a representation of QQ, and μ=σ/τ\mu=\sigma/\tau a slope. There is a deterministic algorithm that finds the HN-filtration of VV with respect to μ\mu in time polynomial in the size of QQ, α(Q0)\alpha(Q_{0}), bit complexity of VV, and absolute values of the entries of σ\sigma and τ\tau.

This result improves a recent result [Che24] which runs in time polynomial in the number of paths in QQ.

Recently, [HS24] introduced the coarse Dulmage-Mendelsohn (DM) decomposition of a linear matrix, generalizing the classic DM-decomposition of a bipartite graph. They showed that a natural gradient flow of operator scaling converges to the coarse DM-decomposition. However, their result did not provide an efficient algorithm to compute the coarse DM-decomposition because the gradient flow may take exponential time to converge. We show that the coarse DM-decomposition is indeed a special case of the HN-filtration for the generalized Kronecker quiver. Since the absolute values of the weights involved for this special case are polynomially bounded, our algorithm finds the coarse DM-decomposition in polynomial time; see Section 4.3.

1.4 King’s polyhedral cone, rank-one representations, and submodular flow

Motivated by King’s criterion, we investigate a polyhedral cone that is the set of σQ0\sigma\in\mathbb{R}^{Q_{0}} satisfying σ(dim¯V)=0\sigma(\operatorname{\underline{dim}}V)=0 and σ(dim¯W)0\sigma(\operatorname{\underline{dim}}W)\leq 0 for any subrepresentation WW of VV. Since the number of distinct dim¯W\operatorname{\underline{dim}}W is finite, the above linear system is also finite and hence defines a polyhedral cone. We call the polyhedral cone the King cone of a quiver representation VV. Interestingly, the King cone is related to the cone of feasible flows in network-flow problems.

Let us first consider the easiest case. If VV is a quiver representation with dimV(i)=1\dim V(i)=1 for all iQ0i\in Q_{0}, then King’s criterion characterizes the existence of a nonnegative flow φ\varphi on the support quiver with the boundary condition φ=σ\partial\varphi=\sigma. To state it more precisely, we introduce notation on flows. Let Q=(Q0,Q1)Q=(Q_{0},Q_{1}) be a quiver (or a directed graph). For a vertex subset XQ0X\subseteq Q_{0}, let Out(X)\operatorname{Out}(X) denote the set of outgoing arcs from XX, i.e., Out(X){a=(i,j):iX,jQ0X}\operatorname{Out}(X)\coloneqq\{a=(i,j):i\in X,\ j\in Q_{0}\setminus X\}. Similarly, let In(X)\operatorname{In}(X) denote the set of incoming arcs to XX. If X={i}X=\{i\}, we abbreviate Out({i})\operatorname{Out}(\{i\}) and In({i})\operatorname{In}(\{i\}) as Out(i)\operatorname{Out}(i) and In(i)\operatorname{In}(i), respectively. If Out(X)=\operatorname{Out}(X)=\emptyset, then XX is called a lower set of QQ. For a flow φQ1\varphi\in\mathbb{R}^{Q_{1}} on QQ, its boundary φQ0\partial\varphi\in\mathbb{R}^{Q_{0}} is defined by φ(i)aOut(i)φ(a)aIn(i)φ(a)\partial\varphi(i)\coloneqq\sum_{a\in\operatorname{Out}(i)}\varphi(a)-\sum_{a^{\prime}\in\operatorname{In}(i)}\varphi(a^{\prime}) for iQ0i\in Q_{0}.

Let us return to the σ\sigma-semistability of a representation VV with dimV(i)=1\dim V(i)=1 for all iQ0i\in Q_{0}. In this case, V(a)V(a)\in\mathbb{C} for each arc aQ1a\in Q_{1}, and a subrepresentation WW of VV can be identified with a vertex subset XQ0X\subseteq Q_{0}. By the definition of a subrepresentation, if iXi\in X and W(a)0W(a)\neq 0 for a=(i,j)Q1a=(i,j)\in Q_{1}, then jXj\in X. This implies that XX is a lower set in the support quiver of VV, namely, the subquiver of QQ whose arcs are aQ1a\in Q_{1} with V(a)0V(a)\neq 0. Therefore, King’s criterion is equivalent to the purely combinatorial condition that σ(Q0)=0\sigma(Q_{0})=0 and σ(X)0\sigma(X)\leq 0 for each lower set XX of QQ, which characterizes the existence of a nonnegative flow φ\varphi on the support quiver with the boundary condition φ=σ\partial\varphi=\sigma by Gale’s theorem [Gal57] (see, e.g., [KV18, Theorem 9.2]).

By generalizing the above observation, we show that if VV is a rank-one representation of QQ, i.e., V(a)V(a) is a rank-one matrix for each aQ1a\in Q_{1}, then

  • King’s criterion can be rephrased as a purely combinatorial condition with respect to the linear matroids arising from the rank-one matrices V(a)V(a) of QQ, and

  • the rephrased condition above can be further viewed as the feasibility condition of a network flow-type problem called submodular flow.

That is, the King cone is representable as the feasibility of a certain instance of the submodular flow problem. This enables us to decide the σ\sigma-semistability for rank-one representations in strongly polynomial time.

Theorem 1.6 (informal version of Theorems 5.8 and 5.9).

Let QQ be an acyclic quiver, VV a rank-one representation of QQ, and σ\sigma a weight. Then, σ\sigma is in the King cone if and only if there is a feasible flow in the instance of submodular flow constructed from VV, σ\sigma. Therefore, using standard submodular flow algorithms, we can decide the σ\sigma-semistability of rank-one representations in strongly polynomial time.

This theorem recovers the following well-known results when applied to the generalized Kronecker quiver and a star quiver.

  • A rank-one linear matrix k=1mxkvkfk\sum_{k=1}^{m}x_{k}v_{k}f_{k} is (nc-)nonsingular (where vkv_{k} is a column vector and fkf_{k} a row vector) if and only if the linear matroids of (fk:k[m])(f_{k}:k\in[m]) and (vk:k[m])(v_{k}:k\in[m]) have a common base [Lov89].

  • If each linear operator Bi=fiB_{i}=f_{i} is of rank-one for i[m]i\in[m] (where fif_{i} is a row vector), the BL polytope coincides with the base polytope of the linear matroid of (fi:i[m])(f_{i}:i\in[m]) [Bar98].

1.5 Semistability of general quivers

Thus far, we have considered the σ\sigma-semistability of acyclic quivers. As a complementary result, we show that the semistability of cyclic quivers under the GL(Q,α)\operatorname{GL}(Q,\alpha)-action can be efficiently reduced to noncommutative polynomial identity testing. In particular, we show that the polynomial can be represented by an algebraic branching program (ABP). This yields a deterministic algorithm for deciding the semistability of general quivers because noncommutative polynomial identity testing for ABP can be conducted in deterministic polynomial time [RS05]; see Section 6. Note that [BFGO+19] devised another deterministic algorithm for the problem with their framework of noncommutative optimization, which is built upon deep results in various areas of mathematics. See also the discussion in related work.

Remark.

After submitting the first version of this paper, an anonymous reviewer pointed out that this result for general quivers was sketched in [Mul17]. See Theorem 10.8 and the last paragraph of Section 10.2 in [Mul17]. At a very high level, our algorithm and Mulmuley’s results follow a similar strategy, although Mulmuley’s result requires several deep algebro-geometric backgrounds. We believe that our proof is more explicit and elementary, and hence, is worthy to be presented here for completeness.

1.6 Our techniques

In this subsection, we outline our techniques.

σ\sigma-semistability.

Our starting point is a reduction of general acyclic quivers to the generalized Kronecker quiver [DM17, Hus21]. We decompose the weight σ=σ+σ\sigma=\sigma^{+}-\sigma^{-}, where σ+(i)max{σ(i),0}\sigma^{+}(i)\coloneqq\max\{\sigma(i),0\} and σ(i)max{σ(i),0}\sigma^{-}(i)\coloneqq\max\{-\sigma(i),0\}. Let Q0+Q_{0}^{+} and Q0Q_{0}^{-} be the sets of vertices ii such that σ(i)>0\sigma(i)>0 and σ(i)<0\sigma(i)<0, respectively. [DM17] showed that the σ\sigma-semistability of a representation VV with the dimension vector α\alpha is equivalent to the nc-nonsingularity of a partitioned linear matrix

A:iQ0+(α(i))σ+(i)iQ0(α(i))σ(i),\displaystyle A:\bigoplus_{i\in Q_{0}^{+}}{\bigl{(}\mathbb{C}^{\alpha(i)}\bigr{)}}^{\oplus\sigma^{+}(i)}\to\bigoplus_{i\in Q_{0}^{-}}{\bigl{(}\mathbb{C}^{\alpha(i)}\bigr{)}}^{\oplus\sigma^{-}(i)}, (1.8)

where the (s,p;t,q)(s,p;t,q)-block (sQ0+s\in Q_{0}^{+}, p[σ+(s)]p\in[\sigma^{+}(s)], tQ0t\in Q_{0}^{-}, q[σ(t)]q\in[\sigma^{-}(t)]) of AA is given by a linear matrix

P:st pathxP,p,qV(P).\sum_{P:\text{$s$--$t$ path}}x_{P,p,q}V(P).

Here, xP,p,qx_{P,p,q} is an indeterminate and V(P)V(P) is the linear map corresponding to the path PP, i.e., V(P)V(ak)V(a1)V(P)\coloneqq V(a_{k})\cdots V(a_{1}) for P=(a1,,ak)P=(a_{1},\dots,a_{k}) as a sequence of arcs. However, because the number of indeterminates is exponential, applying the known nc-rank computation algorithms in a black box manner does not yield the desired time complexity.

Inspired by the above reduction, we introduce the following scaling problem for quiver representations. We define a scaling Vg,hV_{g,h} of the quiver representation VV by the block matrices g=(gt:tQ0)g=\bigoplus(g_{t}:t\in Q_{0}^{-}) and h=(hs:sQ0+)h=\bigoplus(h_{s}:s\in Q_{0}^{+}) as

Vg,h(a){ghaV(a)htaif aOut(Q0+)In(Q0),V(a)htaif aIn(Q0)Out(Q0+),ghaV(a)if aOut(Q0+)In(Q0),V(a)otherwise.\displaystyle V_{g,h}(a)\coloneqq\begin{cases}g_{ha}V(a)h_{ta}^{\dagger}&\text{if $a\in\operatorname{Out}(Q_{0}^{+})\cap\operatorname{In}(Q_{0}^{-})$},\\ V(a)h_{ta}^{\dagger}&\text{if $a\in\operatorname{In}(Q_{0}^{-})\setminus\operatorname{Out}(Q_{0}^{+})$},\\ g_{ha}V(a)&\text{if $a\in\operatorname{Out}(Q_{0}^{+})\setminus\operatorname{In}(Q_{0}^{-})$},\\ V(a)&\text{otherwise}.\end{cases} (1.9)

Furthermore, let (b+,b)iQ0+α(s)×tQ0α(t)(b^{+},b^{-})\in\bigoplus_{i\in Q_{0}^{+}}\mathbb{Q}^{\alpha(s)}\times\bigoplus_{t\in Q_{0}^{-}}\mathbb{Q}^{\alpha(t)} be vectors such that b+(s)=σ+(s)N𝟏α(s)b^{+}(s)=\frac{\sigma^{+}(s)}{N}\mathbf{1}_{\alpha(s)} and b(t)=σ(t)N𝟏α(t)b^{-}(t)=\frac{\sigma^{-}(t)}{N}\mathbf{1}_{\alpha(t)}, where Nσ+(α)=σ(α)N\coloneqq\sigma^{+}(\alpha)=\sigma^{-}(\alpha) and 𝟏\mathbf{1} denotes the all-one vector. We say that VV is approximately scalable (to the marginals (b+,b)(b^{+},b^{-})) if for any ε>0\varepsilon>0, there exist block matrices gg and hh such that Vg,hV_{g,h} satisfies

sQ0+P:st pathVg,h(P)Vg,h(P)Diag(b(t))tr\displaystyle\left\lVert\sum_{s\in Q_{0}^{+}}\sum_{P:\text{$s$--$t$ path}}V_{g,h}(P)V_{g,h}(P)^{\dagger}-\operatorname{Diag}(b^{-}(t))\right\rVert_{\operatorname{tr}} <ε(tQ0),\displaystyle<\varepsilon\qquad(t\in Q_{0}^{-}), (1.10)
tQ0P:st pathVg,h(P)Vg,h(P)Diag(b+(s))tr\displaystyle\left\lVert\sum_{t\in Q_{0}^{-}}\sum_{P:\text{$s$--$t$ path}}V_{g,h}(P)^{\dagger}V_{g,h}(P)-\operatorname{Diag}(b^{+}(s))\right\rVert_{\operatorname{tr}} <ε(sQ0+),\displaystyle<\varepsilon\qquad(s\in Q_{0}^{+}), (1.11)

where the norm is the trace norm.333The choice of the trace norm is not important here; we can use any unitary invariant norm. This is an instance of operator scaling with specified marginals [Fra18, BFGO+18]. The crucial observation is that even though there may exist exponentially many sstt paths, the above matrix sum can be computed efficiently by exploiting the underlying quiver structure. Therefore, we can use a simple iterative algorithm in [BFGO+18] to check VV is approximately scalable for a fixed ε>0\varepsilon>0. We can show that it runs in O(ε2poly(|Q|,α(Q0),b))O(\varepsilon^{-2}\operatorname{poly}(|Q|,\alpha(Q_{0}),b)) time, where bb is the bit complexity of VV. Furthermore, we show that it is sufficient to consider ε=O(1/N)\varepsilon=O(1/N) to decide the σ\sigma-semistability of VV. This yields our algorithm for checking the σ\sigma-semistability of quiver representations.

Maximizers in King’s criterion.

We follow a similar approach to find a maximizer in King’s criterion. We use the above linear matrix of the reduction [DM17] directly and show that the shrunk subspaces of the above linear matrix correspond to the maximizers of King’s criterion. Then, we show that the necessary operations in the recent shrunk subspace algorithm [FSG23] can be performed efficiently, enabling us to find the inclusion-wise minimum maximizer of King’s criterion efficiently. Note that the correspondence between the shrunk subspaces and maximizers of King’s criterion is shown in [Hus21] using abstract algebra; we provide a more direct and elementary proof using submodularity.

HN-filtration.

Our HN-filtration algorithm is based on principal partitions of submodular systems [Fuj09]. For a slope μ=σ/τ\mu=\sigma/\tau, we consider a parametric modular function

fλ(W)λτ(dim¯W)σ(dim¯W)\displaystyle f_{\lambda}(W)\coloneqq\lambda\tau(\operatorname{\underline{dim}}W)-\sigma(\operatorname{\underline{dim}}W) (1.12)

on the subrepresentations WW of VV, where λ\lambda\in\mathbb{R} is a parameter. Let (λ)\mathcal{L}(\lambda) denote the modular lattice of the minimizers of fλf_{\lambda} and let W(λ)W^{-}(\lambda) and W+(λ)W^{+}(\lambda) be the minimum and maximum minimizers of fλf_{\lambda}, respectively. By the standard argument in principal partition, we show that W+(λ)W(λ)W^{+}(\lambda)\leq W^{-}(\lambda^{\prime}) for λ>λ\lambda>\lambda^{\prime}. Furthermore, there must be a finite set of λ\lambda such that (λ)\mathcal{L}(\lambda) consists of more than one element. We call such a value of λ\lambda a critical value. Let λ1>>λk\lambda_{1}>\cdots>\lambda_{k} be the critical values. Then, they induce the filtration

{0}=W(λ1)<W+(λ1)=W(λ2)<W+(λ2)==W(λk)<W+(λk)=V.\displaystyle\{0\}=W^{-}(\lambda_{1})<W^{+}(\lambda_{1})=W^{-}(\lambda_{2})<W^{+}(\lambda_{2})=\cdots=W^{-}(\lambda_{k})<W^{+}(\lambda_{k})=V. (1.13)

We show that this coincides with the HN-filtration. Each W(λ)W^{-}(\lambda) and W+(λ)W^{+}(\lambda) can be found by our algorithm for maximizers of King’s criterion for fixed λ\lambda. The possible candidates for critical values can be easily enumerated, enabling us to find the HN-filtration efficiently.

Strongly polynomial-time algorithm for rank-one representations.

When VV is a rank-one representation, each rank-one matrix V(a)V(a) is representable as vafav_{a}f_{a} for some nonzero vector vaV(ha)v_{a}\in V(ha) and nonzero dual vector faV(ta)f_{a}\in V(ta)^{*}. Based on this representation, we assign each vertex iQ0i\in Q_{0} to two linear matroids 𝐌i+\mathbf{M}_{i}^{+} and 𝐌i\mathbf{M}_{i}^{-}, where the first is generated by {fa:faOut(i)}\{f_{a}:f_{a}\in\operatorname{Out}(i)\} and the second by {va:vaIn(i)}\{v_{a}:v_{a}\in\operatorname{In}(i)\}. Then, we can simulate a subrepresentation WW of VV as a lower set XX of the directed graph D[V]D[V] constructed from VV; its vertex set is the (disjoint) union of the ground sets {fa:faOut(i)}\{f_{a}:f_{a}\in\operatorname{Out}(i)\} and {va:vaIn(i)}\{v_{a}:v_{a}\in\operatorname{In}(i)\} of the matroids 𝐌i+\mathbf{M}_{i}^{+} and 𝐌i\mathbf{M}_{i}^{-} for iQ0i\in Q_{0}; its arc set represents Q1Q_{1} and the dependencies as “if W(ha)=W(tb)W(ha)=W(tb) contains vav_{a} then W(tb)kerfbW(tb)\not\leq\ker f_{b}; hence, W(hb)W(hb) must contain vbv_{b} for WW to be a subrepresentation of VV.” This enables us to rephrase King’s criterion as a combinatorial condition on the lower sets XX of D[V]D[V] as

iQ0(σ+(i)(dimV(i)ri+({fa:faOut(i)}X))σ(i)ri({va:vaIn(i)}X))0,\displaystyle\sum_{i\in Q_{0}}\left(\sigma^{+}(i)\left(\dim V(i)-r_{i}^{+}(\{f_{a}:f_{a}\in\operatorname{Out}(i)\}\setminus X)\right)-\sigma^{-}(i)r_{i}^{-}(\{v_{a}:v_{a}\in\operatorname{In}(i)\}\cap X)\right)\leq 0, (1.14)

where ri+r_{i}^{+} and rir_{i}^{-} denote the rank functions of 𝐌i+\mathbf{M}_{i}^{+} and 𝐌i\mathbf{M}_{i}^{-}, respectively. We further rephrase the above combinatorial condition as the feasibility characterization of a certain instance of the submodular flow problem by Frank [Fra84]. Thus, we can check the σ\sigma-semistability of a rank-one representation VV by checking the feasibility of the instance generated by VV of the submodular flow problem.

Semistability of general quivers.

For the semistability of general quivers, we use an invariant polynomial characterization of the null-cone. By the general theory of GIT, a representation VV is semistable if and only if there exists a GL(Q,α)\operatorname{GL}(Q,\alpha)-invariant homogeneous polynomial pp on the representation space Rep(Q,α)\operatorname{Rep}(Q,\alpha) such that p(V)0p(V)\neq 0. The Le Bruyn-Procesi theorem [BP90] stated that the ring of invariant polynomials is generated by polynomials in the form of

tr[V(ak)V(a2)V(a1)],\displaystyle\operatorname{tr}[V(a_{k})\cdots V(a_{2})V(a_{1})], (1.15)

for a closed path444Here, a closed path means a sequence (a1,,ak)(a_{1},\dots,a_{k}) of arcs such that hal=tal+1ha_{l}=ta_{l+1} (l[k]l\in[k]), where ak+1a1a_{k+1}\coloneqq a_{1}. In graph theory, it is usually called a closed walk. In this paper, we follow the standard terminologies in quiver representation. (a1,a2,,ak)(a_{1},a_{2},\dots,a_{k}) in QQ with length k1k\geq 1. Furthermore, closed paths with length 1kα(Q0)21\leq k\leq\alpha(Q_{0})^{2} generate the invariant ring, where α=dim¯V\alpha=\operatorname{\underline{dim}}V.

Therefore, we can decide the semistability by checking whether the above polynomial is nonzero at some vertex ii and closed path CC. The obstacle is that the number of closed paths can be exponential. To this end, we consider another polynomial in noncommutative indeterminate xax_{a} (aQ1a\in Q_{1}) defined as

C: closed path starting at ixCtrV(C)\displaystyle\sum_{\text{$C$: closed path starting at $i$}}x^{C}\operatorname{tr}V(C) (1.16)

for each vertex iQ0i\in Q_{0}, where xC=xakxa1x^{C}=x_{a_{k}}\cdots x_{a_{1}} for C=(a1,,ak)C=(a_{1},\dots,a_{k}). Then, VV is semistable if and only if this noncommutative polynomial is nonzero at some vertex ii. Note that the noncommutativity is essential to distinguish closed paths with the same arc sets. For example, consider a quiver with a single vertex and two self-loops (say, aa and bb). Then, closed walks C=abababC=abab\cdots ab and C=akbkC^{\prime}=a^{k}b^{k} of length 2k2k have the same number of aa and bb, but their trace trV(C)\operatorname{tr}V(C) and trV(C)\operatorname{tr}V(C^{\prime}) are different in general for k2k\geq 2. So if we use commutative indeterminates, multiple closed walks correspond to a single monomial, and we cannot decide whether trV(C)0\operatorname{tr}V(C)\neq 0 for some CC or not by checking whether the polynomial is zero or not.

Yet, we need to show how to perform noncommutative polynomial identity testing for this polynomial in deterministic polynomial time. Using the underlying quiver structure, we can show that this polynomial can written as an ABP of polynomial size. Applying the algorithm of [RS05], we obtain our algorithm for the semistability of general quivers.

We remark that this is the only place where a nontrivial algebraic result from the GIT machinery is needed. The other algorithms and analysis can be understood with elementary linear algebra (assuming the known analysis of operator scaling algorithms, which involves some abstract algebra).

1.7 Related work

Existing studies on algorithms for quiver semistability have focused on bipartite quivers [CK21, CD21, CI22, FM23]. We remark that semistability in bipartite quivers is essentially operator scaling with a block structure; see Section 2.4. In bipartite quivers, the number of paths is equal to the number of arcs; hence, a weak running time was sufficient in the previous studies. [Hus21, Che24] studied general acyclic quivers. They first used the reduction of [DM17] to the generalized Kronecker quiver and applied the nc-rank algorithm [IQS18] in a black box manner. Hence, their algorithm runs in time polynomial in the number of paths, which can be exponential.

Several polyhedral cones associated with quiver representations have been studied in the literature [CCK22, VW23]. The moment cone of a quiver QQ and a dimension vector α\alpha is the polyhedral cone generated by the highest weights of the representations of QQ with the dimension vector α\alpha. The membership of the moment cone can be decided in strongly polynomial time for bipartite quivers [CCK22] and even general acyclic quivers [VW23]. Another polyhedral cone is the conic hull of weight σQ0\sigma\in\mathbb{Z}^{Q_{0}} such that there exists a nonzero semi-invariant polynomial in Rep(Q,α)\operatorname{Rep}(Q,\alpha) with weight σ\sigma. The membership problem of this cone is called the generic semistability problem [CCK22]. By definition, σ\sigma is in this cone if and only if there exists a generic σ\sigma-semistable representation of QQ, hence the name. To the best of our knowledge, the generic semistability problem remains open for general acyclic quivers.

The semistability of quiver representations is a special case of semistability in the GIT. In the most abstract setting, GIT studies group actions on algebraic varieties. We say that a point in the variety is semistable if its orbit closure does not contain the origin. [BFGO+19] proposed a framework of noncommutative optimization to devise algorithms for GIT problems in the general setting. Although noncommutative optimization is a broad and general framework, it does not always provide efficient algorithms for all GIT problems. Currently, most of the known tractable problems originate from a family of operator scaling problems, which are also contained in the semistability of quiver representations. Furthermore, it is built upon deep results in various areas of mathematics, such as algebraic geometry, Lie algebra, and representation theory, rendering it difficult for non-experts to understand. Another conceptual contribution of this paper is identifying the semistability of quiver representations as a useful subclass of GIT. The semistability of quiver representations is rich enough to capture various interesting problems in the literature while also supporting the design efficient algorithms using elementary techniques.

1.8 Organization of this paper

The rest of this paper is organized as follows. Section 2 introduces the necessary background and notation of quiver representations, operator scaling, and noncommutative rank. Section 3 presents our algorithms for deciding the σ\sigma-semistability and finding maximizers of King’s criterion. Section 4 presents our HN-filtration algorithm and its application to the coarse DM-decomposition. Section 5 investigates the King cone of rank-one representations and shows the reduction to submodular flow. Section 6 describes our reduction from the semistability of general quivers to polynomial identity testing of noncommutative ABPs.

2 Preliminaries

We denote the set of nonnegative integers, rational, and real numbers by +\mathbb{Z}_{+}, +\mathbb{Q}_{+}, and +\mathbb{R}_{+}, respectively. We let [m,n]{m,m+1,,n1,n}[m,n]\coloneqq\{m,m+1,\dots,n-1,n\} for m,nm,n\in\mathbb{Z} with mnm\leq n and [n][1,n]={1,,n}[n]\coloneqq[1,n]=\{1,\dotsc,n\} for a positive integer nn. We denote the set of m×nm\times n complex matrices by Mat(m,n)\operatorname{Mat}(m,n). We simply denote Mat(n,n)\operatorname{Mat}(n,n) by Mat(n)\operatorname{Mat}(n). The conjugate transpose of a matrix AA is denoted by AA^{\dagger}. The subgroup of the upper triangular matrices in GL(n)\operatorname{GL}(n) (i.e., the Borel subgroup) is denoted by B(n)\mathrm{B}(n). For two vector spaces UU and VV, we mean by UVU\leq V that UU is a subspace of VV. Let S\langle S\rangle denote the vector space spanned by a multiset SS of the vectors. For a vector bnb\in\mathbb{C}^{n}, we denote by Diag(b)\operatorname{Diag}(b) the n×nn\times n diagonal matrix such that the entries of bb are on the diagonal.

2.1 Operator scaling, matrix space, and noncommutative rank

A linear map Φ:Mat(n)Mat(m)\Phi:\operatorname{Mat}(n)\to\operatorname{Mat}(m) is said to be completely positive, or CP for short, if Φ(X)==1kAXA\Phi(X)=\sum_{\ell=1}^{k}A_{\ell}XA_{\ell}^{\dagger} for some AMat(m,n)A_{\ell}\in\operatorname{Mat}(m,n). These AA_{\ell} are called the Kraus operators of Φ\Phi. The dual map Φ:Mat(m)Mat(n)\Phi^{*}:\operatorname{Mat}(m)\to\operatorname{Mat}(n) of Φ\Phi is defined by Φ(X)==1kAXA\Phi^{*}(X)=\sum_{\ell=1}^{k}A_{\ell}^{\dagger}XA_{\ell}. For (g,h)GL(m)×GL(n)(g,h)\in\operatorname{GL}(m)\times\operatorname{GL}(n), we define the scaling Φg,h\Phi_{g,h} of Φ\Phi by

Φg,h(X)gΦ(hXh)g==1k(gAh)X(gAh).\displaystyle\Phi_{g,h}(X)\coloneqq g\Phi(h^{\dagger}Xh)g^{\dagger}=\sum_{\ell=1}^{k}(gA_{\ell}h^{\dagger})X(gA_{\ell}h^{\dagger})^{\dagger}. (2.1)

If m=nm=n, the CP map is said to be square.

Let Φ:Mat(n)Mat(n)\Phi:\operatorname{Mat}(n)\to\operatorname{Mat}(n) be a square CP map. Let 𝖽𝗌(Φ)Φ(I)IF2+Φ(I)IF2\operatorname{\mathsf{ds}}(\Phi)\coloneqq\lVert\Phi(I)-I\rVert_{\mathrm{F}}^{2}+\lVert\Phi^{*}(I)-I\rVert_{\mathrm{F}}^{2}, where F\lVert\cdot\rVert_{\mathrm{F}} denotes the Frobenius norm. Then, Φ\Phi is said to be approximately scalable if for any ε>0\varepsilon>0, there exists (g,h)GL(n)×GL(n)(g,h)\in\operatorname{GL}(n)\times\operatorname{GL}(n) such that 𝖽𝗌(Φg,h)ε\operatorname{\mathsf{ds}}(\Phi_{g,h})\leq\varepsilon. The goal of the operator scaling problem is to decide whether a given CP map is approximately scalable or not.

Operator scaling is closely related to the noncommutative rank (nc-rank) of linear matrices. An m×nm\times n symbolic matrix AA is called a linear matrix if A==1kxAA=\sum_{\ell=1}^{k}x_{\ell}A_{\ell} for indeterminates xx_{\ell} and matrices AMat(m,n)A_{\ell}\in\operatorname{Mat}(m,n). Sometimes, it is more convenient to see a linear matrix as a matrix space spanned by A1,,AkA_{1},\dots,A_{k}. We denote the corresponding matrix space of a linear matrix AA by 𝒜\mathcal{A}. For a subspace UnU\leq\mathbb{C}^{n}, let 𝒜U{Au:A𝒜,uU},\mathcal{A}U\coloneqq\langle\{Au:A\in\mathcal{A},\ u\in U\}\rangle, which is a subspace of m\mathbb{C}^{m}. The nc-rank of a linear matrix AA (denoted by ncrankA\operatorname{nc-rank}A) is defined by

ncrankAmin{n+dim𝒜UdimU:Un}.\displaystyle\operatorname{nc-rank}A\coloneqq\min\{n+\dim\mathcal{A}U-\dim U:U\leq\mathbb{C}^{n}\}. (2.2)

A square linear matrix AA is said to be nc-nonsingular if ncrankA=n\operatorname{nc-rank}A=n. Informally speaking, ncrankA\operatorname{nc-rank}A is the rank of AA, where the indeterminates xix_{i} are pairwise noncommutative, i.e., xixjxjxix_{i}x_{j}\neq x_{j}x_{i} for iji\neq j. See [Coh95, FR04] for more details. A pair (L,R)(L,R) of subspaces LmL\leq\mathbb{C}^{m} and RnR\leq\mathbb{C}^{n} is called an independent subspace if L𝒜R={0}L\cap\mathcal{A}R=\{0\}. Over the complex field, (L,R)(L,R) is independent if and only if tr(ΠLΦ(ΠR))=0\operatorname{tr}(\Pi_{L}\Phi(\Pi_{R}))=0, where ΠL\Pi_{L} denotes the orthogonal projection matrix onto LL. Then,

ncrankA=m+nmax{dimL+dimR:(L,R) an independent subspace}.\displaystyle\operatorname{nc-rank}A=m+n-\max\{\dim L+\dim R:\text{$(L,R)$ an independent subspace}\}. (2.3)

An independent subspace (L,R)(L,R) is said to be maximum if dimL+dimR\dim L+\dim R is maximum. In particular, a square linear matrix AA is nc-nonsingular if and only if dimL+dimRn\dim L+\dim R\leq n for any independent subspace (L,R)(L,R). Gurvits’ theorem [Gur04] states that a square CP map Φ\Phi with the Kraus operator AA_{\ell} ([k]\ell\in[k]) is approximately scalable if and only if the linear matrix =1kxA\sum_{\ell=1}^{k}x_{\ell}A_{\ell} is nc-nonsingular.

2.2 Finding minimum shrunk subspace

Shrunk subspaces are minimizers of the objective function in nc-rank:

f(U)=n+dim𝒜UdimU.\displaystyle f(U)=n+\dim\mathcal{A}U-\dim U. (2.4)

It is easy to see that ff is submodular, i.e., f(U)+f(V)f(UV)+f(U+V)f(U)+f(V)\geq f(U\cap V)+f(U+V) for any subspaces U,VnU,V\leq\mathbb{C}^{n}. Therefore, the shrunk subspaces form a sublattice of n\mathbb{C}^{n}, i.e., if UU and VV are shrunk subspaces, then UVU\cap V and U+VU+V are also shrunk subspaces. Hence, there exists a unique shrunk subspace with minimum dimension, which we call the minimum shrunk subspace. A shrunk subspace can be regarded as a certificate of the value of the nc-rank. In the GIT perspective, shrunk subspaces correspond to one-parameter subgroups that bring the matrix tuple (A1,,Ak)(A_{1},\dots,A_{k}) to the origin. The minimum shrunk subspace is particularly important because it has a rational basis with polynomial bit complexity if (A1,,Ak)(A_{1},\dots,A_{k}) has Gaussian integer entries [IQS18].

Several algorithms can find the minimum shrunk subspace in polynomial time [IQS18, FSG23]. We use the algorithm in [FSG23], which is based on a modified operator Sinkhorn iteration. We use it as a black box and consider only the square case for simplicity. The details of the algorithm can be referred from [FSG23].

Theorem 2.1 ([FSG23]).

Let Φ:Mat(n)Mat(n)\Phi:\operatorname{Mat}(n)\to\operatorname{Mat}(n) be a square CP map whose Kraus operators have Gaussian integer entries. Let AA be the linear matrix corresponding to Φ\Phi. Then, there exists a deterministic polynomial time algorithm that computes a basis of the minimum shrunk subspace of AA. Furthermore, this algorithm works even if Φ\Phi is given as an oracle that computes Φ(X)\Phi(X) and Φ(X)\Phi^{*}(X) for XMat(n)X\in\operatorname{Mat}(n), along with an upper bound bb of the bit complexity of the Kraus operators. The time complexity is polynomial in poly(n,b)(EO+O(n3))\operatorname{poly}(n,b)(\mathrm{EO}+O(n^{3})), where EO\mathrm{EO} denotes the time complexity of a single oracle call.

2.3 Operator scaling with specified marginals

Operator scaling with specified marginals is a generalization of operator scaling. Let (b+,b)+n×+m(b^{+},b^{-})\in\mathbb{R}^{n}_{+}\times\mathbb{R}^{m}_{+} be a pair of nonincreasing nonnegative vectors, which we call the target marginals. We say that a CP map Φ:Mat(n)Mat(m)\Phi:\operatorname{Mat}(n)\to\operatorname{Mat}(m) is approximately scalable to the target marginals (b+,b)(b^{+},b^{-}) if there exist nonsingular upper triangular matrices (g,h)B(m)×B(n)(g,h)\in\mathrm{B}(m)\times\mathrm{B}(n) such that

Φg,h(I)Diag(b)trε,Φg,h(I)Diag(b+)trε.\displaystyle\lVert\Phi_{g,h}(I)-\operatorname{Diag}(b^{-})\rVert_{\operatorname{tr}}\leq\varepsilon,\quad\lVert\Phi_{g,h}^{*}(I)-\operatorname{Diag}(b^{+})\rVert_{\operatorname{tr}}\leq\varepsilon. (2.5)

Such target marginals are said to be feasible. We define Δb++n\Delta b^{+}\in\mathbb{R}_{+}^{n} as Δbj+=bj+bj+1+\Delta b^{+}_{j}=b^{+}_{j}-b^{+}_{j+1} for j[n]j\in[n], where we conventionally define bn+1+0b^{+}_{n+1}\coloneqq 0. Similarly, we define Δb+m\Delta b^{-}\in\mathbb{R}_{+}^{m}. Let Fj+=𝐞1,,𝐞jF^{+}_{j}=\langle\mathbf{e}_{1},\dots,\mathbf{e}_{j}\rangle be the standard flag of n\mathbb{C}^{n} for j[n]j\in[n]. Similarly, we define FiF^{-}_{i} for i[m]i\in[m]. The following theorem characterizes the set of feasible marginals by a certain linear system.

Theorem 2.2 ([Fra18, Theorem 18]).

Let Φ:Mat(n)Mat(m)\Phi:\operatorname{Mat}(n)\to\operatorname{Mat}(m) be a CP map and (b+,b)+n×+m(b^{+},b^{-})\in\mathbb{R}^{n}_{+}\times\mathbb{R}^{m}_{+} a pair of nonincreasing nonnegative vectors. Then, Φ\Phi is approximately scalable to the marginals (b+,b)(b^{+},b^{-}) if and only if j=1nbj+=i=1mbiB\sum_{j=1}^{n}b^{+}_{j}=\sum_{i=1}^{m}b^{-}_{i}\eqqcolon B and

i=1mΔbidim(LFi)+j=1nΔbj+dim(RFj+)B\displaystyle\sum_{i=1}^{m}\Delta b^{-}_{i}\dim(L\cap F^{-}_{i})+\sum_{j=1}^{n}\Delta b^{+}_{j}\dim(R\cap F^{+}_{j})\leq B (2.6)

for any independent subspace (L,R)(L,R) of Φ\Phi.

Let (b+,b)(b^{+},b^{-}) be a feasible marginal with rational entries. There is an efficient algorithm that finds a scaling of a given CP map Φ\Phi whose marginal is ε\varepsilon-close to (b+,b)(b^{+},b^{-}) [Fra18, BFGO+18].

Theorem 2.3 (Theorem 1.13 in [BFGO+18], specialized for operator scaling).

Let ε>0\varepsilon>0 be an accuracy parameter and Φ:Mat(n)Mat(m)\Phi:\operatorname{Mat}(n)\to\operatorname{Mat}(m) a CP map with Gaussian integer Kraus operators. Let (b+,b)n×m(b^{+},b^{-})\in\mathbb{Q}^{n}\times\mathbb{Q}^{m} be target marginals such that b1+bn+0b^{+}_{1}\geq\dots\geq b^{+}_{n}\geq 0, b1bm0b^{-}_{1}\geq\dots\geq b^{-}_{m}\geq 0, and j=1nbj+=i=1mbi=1\sum_{j=1}^{n}b^{+}_{j}=\sum_{i=1}^{m}b^{-}_{i}=1. Then, Algorithm 1 finds upper triangular g,hg,h such that Φg,h(I)Diag(b)trε\lVert\Phi_{g,h}(I)-\operatorname{Diag}(b^{-})\rVert_{\operatorname{tr}}\leq\varepsilon and Φg,h(I)Diag(b+)trε\lVert\Phi_{g,h}^{*}(I)-\operatorname{Diag}(b^{+})\rVert_{\operatorname{tr}}\leq\varepsilon in T=O(ε2(b+Nlog(N)))T=O(\varepsilon^{-2}(b+N\log(\ell N))) iterations, where bb is the maximum bit length of the target marginals (b+,b)(b^{+},b^{-}), Nmax{m,n}N\coloneqq\max\{m,n\}, and \ell is the smallest positive integer such that (b+,b)\ell(b^{+},b^{-}) is an integer. Furthermore, each iteration can be executed in time O(N3)O(N^{3}).

Algorithm 1 Operator Sinkhorn iteration for specified marginals [Fra18, BFGO+18]
1:for t=1,,Tt=1,\dots,T do
2:     \trianglerightLeft normalization
3:     Compute the Cholesky decomposition CC=Φ(I)CC^{\dagger}=\Phi(I). Set g=Diag(b)1/2C1g=\operatorname{Diag}(b^{-})^{1/2}C^{-1} and ΦΦg,I\Phi\leftarrow\Phi_{g,I}.
4:     \trianglerightRight normalization
5:     Compute the Cholesky decomposition CC=Φ(I)CC^{\dagger}=\Phi^{*}(I). Set h=Diag(b+)1/2C1h=\operatorname{Diag}(b^{+})^{1/2}C^{-1} and ΦΦI,h\Phi\leftarrow\Phi_{I,h}.

2.4 Useful results for block matrices

We frequently use CP maps or linear matrices with block structures throughout the paper. Hence we present some useful results here. Let V+V^{+} and VV^{-} be finite sets and let α(s)\alpha(s) (sV+s\in V^{+}) and α(t)\alpha(t) (tVt\in V^{-}) be positive integers. Let nsV+α(s)n\coloneqq\sum_{s\in V^{+}}\alpha(s) and mtVα(t)m\coloneqq\sum_{t\in V^{-}}\alpha(t). Consider a linear matrix AA of size m×nm\times n with the following block structure: The (s,t)(s,t)-block of AA is given by a linear matrix

=1ks,txs,t,As,t,,\displaystyle\sum_{\ell=1}^{k_{s,t}}x_{s,t,\ell}A_{s,t,\ell}, (2.7)

where ks,tk_{s,t} is a nonnegative integer, xs,t,x_{s,t,\ell} is an indeterminate, and As,t,A_{s,t,\ell} is an α(t)×α(s)\alpha(t)\times\alpha(s) matrix. We remark that the indeterminate xs,t,x_{s,t,\ell} appears only in the (s,t)(s,t)-block of AA. Let 𝒜s,t\mathcal{A}_{s,t} denote the matrix subspace spanned by As,t,A_{s,t,\ell}.

The following is a useful lemma that shows the existence of a shrunk subspace respecting the block structure.

Lemma 2.4.

For a partitioned linear matrix in the form (2.7), every shrunk subspace UU is in the form U=sV+UsU=\bigoplus_{s\in V^{+}}U_{s}, where Usα(s)U_{s}\leq\mathbb{C}^{\alpha(s)}.

Proof.

For any UnU\leq\mathbb{C}^{n}, let UsU_{s} denote the projection of UU to α(s)\mathbb{C}^{\alpha(s)} for sV+s\in V^{+}. Then, we have

𝒜U=tVsV+𝒜s,tUs.\mathcal{A}U=\bigoplus_{t\in V^{-}}\sum_{s\in V^{+}}\mathcal{A}_{s,t}U_{s}.

Hence, replacing UU with UsV+UsU^{\prime}\coloneqq\bigoplus_{s\in V^{+}}U_{s} does not change 𝒜U\mathcal{A}U. If UUU\neq U^{\prime}, then dimU<dimU\dim U<\dim U^{\prime}. Hence, we have dim𝒜UdimU>dim𝒜UdimU\dim\mathcal{A}U-\dim U>\dim\mathcal{A}U^{\prime}-\dim U^{\prime}, which implies that UU is not a shrunk subspace. ∎

The following is deduced from the modular lattice structure of the optimal shrunk subspaces.

Lemma 2.5.

Let AA be a partitioned linear matrix as in (2.7) and U=sV+UsU=\bigoplus_{s\in V^{+}}U_{s} the minimal shrunk subspace of AA. Suppose that there exist s,sV+s,s^{\prime}\in V^{+} such that 𝒜s,t=𝒜s,t\mathcal{A}_{s,t}=\mathcal{A}_{s^{\prime},t} for all tVt\in V^{-}. Then, Us=UsU_{s}=U_{s^{\prime}}.

Let us consider the corresponding CP map Φ:sV+Mat(ns)tVMat(nt)\Phi:\bigoplus_{s\in V^{+}}\operatorname{Mat}(n_{s})\to\bigoplus_{t\in V^{-}}\operatorname{Mat}(n_{t}) that maps X=sV+XsX=\bigoplus_{s\in V^{+}}X_{s} to Φ(X)=tVΦ(X)\Phi(X)=\bigoplus_{t\in V^{-}}\Phi(X), where

Φ(X)t=sV+=1ks,tAs,t,XsAs,t,.\displaystyle\Phi(X)_{t}=\sum_{s\in V^{+}}\sum_{\ell=1}^{k_{s,t}}A_{s,t,\ell}X_{s}A_{s,t,\ell}^{\dagger}. (2.8)

Here is a version of Theorem 2.2 for CP maps with a block structure. We say that a matrix hsV+Mat(ns)h\in\bigoplus_{s\in V^{+}}\operatorname{Mat}(n_{s}) is block-wise upper triangular if h=sV+hsh=\bigoplus_{s\in V^{+}}h_{s} where hsh_{s} is upper triangular for all sV+s\in V^{+}.

Lemma 2.6 ([Fra18, Proposition 61]).

Let Φ\Phi be a CP map in the form (2.7). Let (b+,b)=(sV+b+(s),tVb(s))+n×+m(b^{+},b^{-})=(\bigoplus_{s\in V^{+}}b^{+}(s),\bigoplus_{t\in V^{-}}b^{-}(s))\in\mathbb{R}^{n}_{+}\times\mathbb{R}^{m}_{+} be a pair of nonnegative vectors such that b+(s)+α(s)b^{+}(s)\in\mathbb{R}_{+}^{\alpha(s)} and b(t)+α(t)b^{-}(t)\in\mathbb{R}_{+}^{\alpha(t)} are nonincreasing for sV+s\in V^{+} and tVt\in V^{-}. Then, Φ\Phi is approximately scalable to the marginals (b+,b)(b^{+},b^{-}) if and only if sV+j=1α(s)b+(s)j=tVi=1α(t)b(t)iB\sum_{s\in V^{+}}\sum_{j=1}^{\alpha(s)}b^{+}(s)_{j}=\sum_{t\in V^{-}}\sum_{i=1}^{\alpha(t)}b^{-}(t)_{i}\eqqcolon B and

tVi=1α(t)Δb(t)idim(LtF(t)i)+sV+j=1α(s)Δb+(s)jdim(RsF+(s)j)B\displaystyle\sum_{t\in V^{-}}\sum_{i=1}^{\alpha(t)}\Delta b^{-}(t)_{i}\dim(L_{t}\cap F^{-}(t)_{i})+\sum_{s\in V^{+}}\sum_{j=1}^{\alpha(s)}\Delta b^{+}(s)_{j}\dim(R_{s}\cap F^{+}(s)_{j})\leq B (2.9)

for any independent subspace (L,R)=(tVLt,sV+Rs)(L,R)=(\bigoplus_{t\in V^{-}}L_{t},\bigoplus_{s\in V^{+}}R_{s}). Furthermore, scaling matrices can be taken to be block-wise upper triangular.

The target marginal often has the same block structure. We say that the target marginal (b+,b)(b^{+},b^{-}) respects the block structure if b+b^{+} and bb^{-} are constant within each block.

Lemma 2.7.

Let Φ\Phi be a CP map with a block structure as in (2.8) and (b+,b)(b^{+},b^{-}) a target marginal respecting the same block structure. Let ε>0\varepsilon>0 be an accuracy parameter. Then, Φ\Phi can be scaled by block-wise upper triangular matrices g,hg,h such that Φg,h(I)Diag(b)trε\lVert\Phi_{g,h}(I)-\operatorname{Diag}(b^{-})\rVert_{\operatorname{tr}}\leq\varepsilon and Φg,h(I)Diag(b+)trε\lVert\Phi_{g,h}^{*}(I)-\operatorname{Diag}(b^{+})\rVert_{\operatorname{tr}}\leq\varepsilon if and only if the same is possible with block nonsingular matrices g,hg,h.

Proof.

By abusing the notation, we denote the common value of the entries of the ssth block of b+b^{+} by b+(s)b^{+}(s); the same is used for b(t)b^{-}(t) as well. First, observe that for block unitary matrices g,hg,h

Φg,h(I)tDiag(b)t\displaystyle\Phi_{g,h}(I)_{t}-\operatorname{Diag}(b^{-})_{t} =gtΦ(hh)tgtb(t)Iα(t)=gt(Φ(I)tDiag(b)t)gt,\displaystyle=g_{t}\Phi(h^{\dagger}h)_{t}g_{t}^{\dagger}-b^{-}(t)I_{\alpha(t)}=g_{t}(\Phi(I)_{t}-\operatorname{Diag}(b^{-})_{t})g_{t}^{\dagger}, (2.10)
Φg,h(I)sDiag(b+)s\displaystyle\Phi^{*}_{g,h}(I)_{s}-\operatorname{Diag}(b^{+})_{s} =hsΦ(gg)shsb+(s)Iα(s)=hs(Φ(I)sDiag(b+)s)hs\displaystyle=h_{s}\Phi^{*}(g^{\dagger}g)_{s}h_{s}^{\dagger}-b^{+}(s)I_{\alpha(s)}=h_{s}(\Phi^{*}(I)_{s}-\operatorname{Diag}(b^{+})_{s})h_{s}^{\dagger} (2.11)

for sV+s\in V^{+} and tVt\in V^{-}. Since the trace norm is unitary invariant, the approximate scalability condition (i.e., Φg,h(I)Diag(b)trε\lVert\Phi_{g,h}(I)-\operatorname{Diag}(b^{-})\rVert_{\operatorname{tr}}\leq\varepsilon and Φg,h(I)Diag(b+)trε\lVert\Phi_{g,h}^{*}(I)-\operatorname{Diag}(b^{+})\rVert_{\operatorname{tr}}\leq\varepsilon) does not change under the (left) multiplication of block unitary matrices to g,hg,h. Therefore, if there exist nonsingular block matrices g,hg,h that satisfy the approximate scalability condition, the R-factors of the QR decompositions of g,hg,h satisfy the same condition. Therefore, one can take block-wise upper triangular g,hg,h. The other direction is trivial. ∎

3 Reduction to nc-rank and algorithms for semistability

In this section, we present our algorithms for deciding σ\sigma-semistability and finding maximizers of King’s criterion.

3.1 Reduction from semistability to nc-rank

Here, we recall the reduction of the σ\sigma-semistability of an acyclic quiver to nc-nonsingularity testing [DM17].

Let QQ be an acyclic quiver and VV a representation of QQ with the dimension vector α\alpha. Let σQ0\sigma\in\mathbb{Z}^{Q_{0}} be a weight.555We do not assume σ(α)=0\sigma(\alpha)=0 here because the reduction does not need it. This is useful for finding a maximizer in King’s criterion. Let Q0+Q_{0}^{+} and Q0Q_{0}^{-} be the sets of vertices ii such that σ(i)>0\sigma(i)>0 and σ(i)<0\sigma(i)<0, respectively. Let σ+(i)max{σ(i),0}\sigma^{+}(i)\coloneqq\max\{\sigma(i),0\} and σ(i)max{σ(i),0}\sigma^{-}(i)\coloneqq\max\{-\sigma(i),0\} for each iQ0i\in Q_{0}. Note that σ=σ+σ\sigma=\sigma^{+}-\sigma^{-}. Let Nσ+(α)=σ(α)N\coloneqq\sigma^{+}(\alpha)=\sigma^{-}(\alpha). We define an N×NN\times N partitioned linear matrix AA as follows. As a linear map,

A:sQ0+(α(s))σ+(s)tQ0(α(t))σ(t).\displaystyle A:\bigoplus_{s\in Q_{0}^{+}}{\bigl{(}\mathbb{C}^{\alpha(s)}\bigr{)}}^{\oplus\sigma^{+}(s)}\to\bigoplus_{t\in Q_{0}^{-}}{\bigl{(}\mathbb{C}^{\alpha(t)}\bigr{)}}^{\oplus\sigma^{-}(t)}. (3.1)

The (s,p;t,q)(s,p;t,q)-block (sQ0+s\in Q_{0}^{+}, p[σ+(s)]p\in[\sigma^{+}(s)], tQ0t\in Q_{0}^{-}, q[σ(t)]q\in[\sigma^{-}(t)]) of AA is given by a linear matrix

P:st pathxP,p,qV(P),\sum_{P:\text{$s$--$t$ path}}x_{P,p,q}V(P),

where xP,p,qx_{P,p,q} is an indeterminate and V(P)V(P) is the linear map corresponding to the path PP, i.e., V(P)V(ak)V(a1)V(P)\coloneqq V(a_{k})\cdots V(a_{1}) for P=(a1,,ak)P=(a_{1},\dots,a_{k}) as a sequence of arcs. The number of indeterminates in AA is equal to

sQ0+tQ0σ+(s)σ(t)m(s,t),\displaystyle\sum_{s\in Q_{0}^{+}}\sum_{t\in Q_{0}^{-}}\sigma^{+}(s)\sigma^{-}(t)m(s,t), (3.2)

where m(s,t)m(s,t) denotes the number of sstt paths in QQ. Thus, the number of indeterminates may be exponential in general.

The following lemma connects the nc-rank of AA with King’s criterion, which is shown in [Hus21, Theorem 3.3] using abstract algebra. We present an elementary proof for completeness.

Lemma 3.1.

The minimal maximizer of σ(dim¯W)\sigma(\operatorname{\underline{dim}}W) for the subrepresentations WW of VV corresponds to the minimal shrunk subspace of AA.

Proof.

Let WW be a subrepresentation of VV. Define a subspace

UiQ0+W(i)σ+(i)iQ0+(α(i))σ+(i).\displaystyle U\coloneqq\bigoplus_{i\in Q_{0}^{+}}W(i)^{\oplus\sigma^{+}(i)}\leq\bigoplus_{i\in Q_{0}^{+}}(\mathbb{C}^{\alpha(i)})^{\oplus\sigma^{+}(i)}. (3.3)

Inductively, we can show that 𝒜UiQ0W(i)σ(i)\mathcal{A}U\leq\bigoplus_{i\in Q_{0}^{-}}W(i)^{\oplus\sigma^{-}(i)} since WW is a subrepresentation. Hence,

dimUdim𝒜UiQ0+σ+(i)dimW(i)iQ0σ(i)dimW(i)=σ(dim¯W).\displaystyle\dim U-\dim\mathcal{A}U\geq\sum_{i\in Q_{0}^{+}}\sigma^{+}(i)\dim W(i)-\sum_{i\in Q_{0}^{-}}\sigma^{-}(i)\dim W(i)=\sigma(\operatorname{\underline{dim}}W). (3.4)

On the other hand, let UNU\leq\mathbb{C}^{N} be the minimal subspace that maximizes dimUdim𝒜U\dim U-\dim\mathcal{A}U. By the block structure of AA and Lemmas 2.4 and 2.5, we can decompose UU as U=iQ0+Uiσ+(i)U=\bigoplus_{i\in Q_{0}^{+}}U_{i}^{\oplus\sigma^{+}(i)} for some Uiα(i)U_{i}\leq\mathbb{C}^{\alpha(i)}. Define a subrepresentation WW recursively in a topological order of QQ as follows. If iQ0+i\in Q_{0}^{+}, define W(i)=UiW(i)=U_{i}. Otherwise, define W(i)=aIn(i)V(a)W(ta)W(i)=\sum_{a\in\operatorname{In}(i)}V(a)W(ta). We show that WW is indeed a subrepresentation of VV. The only nontrivial case is when there exists a vertex iQ0+i\in Q_{0}^{+} such that aIn(i)V(a)W(ta)Ui\sum_{a\in\operatorname{In}(i)}V(a)W(ta)\not\leq U_{i}. Suppose that this happens. Denote aIn(i)V(a)W(ta)\sum_{a\in\operatorname{In}(i)}V(a)W(ta) by W(i)W^{-}(i). Then, let Ui=Ui+W(i)U^{\prime}_{i}=U_{i}+W^{-}(i) and Uj=UjU^{\prime}_{j}=U_{j} for the rest. This gives a subspace UU^{\prime} that has a strictly larger value of dimUdim𝒜U\dim U^{\prime}-\dim\mathcal{A}U^{\prime} because 𝒜U=𝒜U\mathcal{A}U^{\prime}=\mathcal{A}U, contradicting the assumption of UU. Thus, WW is a subrepresentation of VV. By construction, we have 𝒜U=iQ0W(i)σ(i)\mathcal{A}U=\bigoplus_{i\in Q_{0}^{-}}W(i)^{\oplus\sigma^{-}(i)}. Therefore,

dimUdim𝒜U=iQ0+σ+(i)dimW(i)iQ0σ(i)dimW(i)=σ(dim¯W)\displaystyle\dim U-\dim\mathcal{A}U=\sum_{i\in Q_{0}^{+}}\sigma^{+}(i)\dim W(i)-\sum_{i\in Q_{0}^{-}}\sigma^{-}(i)\dim W(i)=\sigma(\operatorname{\underline{dim}}W) (3.5)

holds as required. ∎

By the lemma, one can check the σ\sigma-semistability of a quiver representation by checking whether the corresponding linear matrix is nc-nonsingular or not. However, the naive reduction does not give a polynomial time algorithm, as the number of indeterminates in the linear matrix may be exponential.

3.2 Scaling algorithm for σ\sigma-semistability

We present a scaling algorithm for deciding the σ\sigma-semistability. The idea is to reduce the problem to operator scaling with specified marginals.

Let us define a CP map ΦV\Phi_{V} corresponding to a quiver representation VV. As a linear map,

ΦV:sQ0+Mat(α(s))tQ0Mat(α(t)).\displaystyle\Phi_{V}:\bigoplus_{s\in Q_{0}^{+}}\operatorname{Mat}(\alpha(s))\to\bigoplus_{t\in Q_{0}^{-}}\operatorname{Mat}(\alpha(t)). (3.6)

Let X=(Xs:sQ0+)X=\bigoplus(X_{s}:s\in Q_{0}^{+}) be an input block matrix. Then,

(ΦV(X))tsQ0+P:st pathV(P)XsV(P)\displaystyle(\Phi_{V}(X))_{t}\coloneqq\sum_{s\in Q_{0}^{+}}\sum_{P:\text{$s$--$t$ path}}V(P)X_{s}V(P)^{\dagger} (3.7)

for tQ0t\in Q_{0}^{-}. The dual map

ΦV:tQ0Mat(α(t))sQ0+Mat(α(s))\displaystyle\Phi_{V}^{*}:\bigoplus_{t\in Q_{0}^{-}}\operatorname{Mat}(\alpha(t))\to\bigoplus_{s\in Q_{0}^{+}}\operatorname{Mat}(\alpha(s)) (3.8)

is given by

(ΦV(Y))stQ0P:st pathV(P)YtV(P)\displaystyle(\Phi_{V}^{*}(Y))_{s}\coloneqq\sum_{t\in Q_{0}^{-}}\sum_{P:\text{$s$--$t$ path}}V(P)^{\dagger}Y_{t}V(P) (3.9)

for sQ0+s\in Q_{0}^{+}. Analogous to the scaling of CP maps, we define a scaling Vg,hV_{g,h} of the quiver representation VV by block matrices g=(gt:tQ0)g=(g_{t}:t\in Q_{0}^{-}) and h=(hs:sQ0+)h=(h_{s}:s\in Q_{0}^{+}) as

Vg,h(a){ghaV(a)htaif aOut(Q0+)In(Q0),V(a)htaif aIn(Q0)Out(Q0+),ghaV(a)if aOut(Q0+)In(Q0),V(a)otherwise.\displaystyle V_{g,h}(a)\coloneqq\begin{cases}g_{ha}V(a)h_{ta}^{\dagger}&\text{if $a\in\operatorname{Out}(Q_{0}^{+})\cap\operatorname{In}(Q_{0}^{-})$},\\ V(a)h_{ta}^{\dagger}&\text{if $a\in\operatorname{In}(Q_{0}^{-})\setminus\operatorname{Out}(Q_{0}^{+})$},\\ g_{ha}V(a)&\text{if $a\in\operatorname{Out}(Q_{0}^{+})\setminus\operatorname{In}(Q_{0}^{-})$},\\ V(a)&\text{otherwise}.\end{cases} (3.10)

Here is the key lemma that relates the σ\sigma-semistability of a quiver representation to the feasibility of a specific marginal of ΦV\Phi_{V}.

Lemma 3.2.

Let VV be a representation of an acyclic quiver QQ with the dimension vector α\alpha and σ\sigma a weight with σ(α)=0\sigma(\alpha)=0. Let (b+,b)(b^{+},b^{-}) be the target marginals such that b+(s)=σ+(s)N𝟏α(s)b^{+}(s)=\frac{\sigma^{+}(s)}{N}\mathbf{1}_{\alpha(s)} and b(t)=σ(t)N𝟏α(t)b^{-}(t)=\frac{\sigma^{-}(t)}{N}\mathbf{1}_{\alpha(t)}, where Nσ+(α)=σ(α)N\coloneqq\sigma^{+}(\alpha)=\sigma^{-}(\alpha). Then, ΦV\Phi_{V} is approximately scalable to the marginals (b+,b)(b^{+},b^{-}) if and only if VV is σ\sigma-semistable.

Proof.

We show that the conditions for scalability in Lemma 2.6 is equivalent to AA being nc-nonsingular, which in turn is equivalent to VV being σ\sigma-semistable. By construction,

Δb+(s)i={σ+(s)Nif i=α(s),0otherwise,Δb(t)j={σ(t)Nif j=α(t),0otherwise.\displaystyle\Delta b^{+}(s)_{i}=\begin{cases}\frac{\sigma^{+}(s)}{N}&\text{if $i=\alpha(s)$},\\ 0&\text{otherwise},\end{cases}\qquad\Delta b^{-}(t)_{j}=\begin{cases}\frac{\sigma^{-}(t)}{N}&\text{if $j=\alpha(t)$},\\ 0&\text{otherwise}.\end{cases} (3.11)

Thus, the condition in Lemma 2.6 reduces to

tQ0σ(t)dimLt+sQ0+σ+(s)dimRsN\displaystyle\sum_{t\in Q_{0}^{-}}\sigma^{-}(t)\dim L_{t}+\sum_{s\in Q_{0}^{+}}\sigma^{+}(s)\dim R_{s}\leq N (3.12)

for any independent subspace (L,R)(L,R). This shows the nc-nonsingularity of AA by the formula of nc-rank in terms of independent subspaces (2.3). ∎

To check the feasibility of the marginal, one can use the scaling algorithm for operator scaling with the specified marginals (Algorithm 1). This yields Algorithm 2.

Algorithm 2 Scaling algorithm for σ\sigma-semistability
1:a representation VV of an acyclic quiver QQ and a weight σ\sigma
2:Let b+(s)σ+(s)N𝟏α(s)b^{+}(s)\coloneqq\frac{\sigma^{+}(s)}{N}\mathbf{1}_{\alpha(s)} and b(t)σ(t)N𝟏α(t)b^{-}(t)\coloneqq\frac{\sigma^{-}(t)}{N}\mathbf{1}_{\alpha(t)}, where N=max{σ+(α),σ(α)}N=\max\{\sigma^{+}(\alpha),\sigma^{-}(\alpha)\}.
3:Set ε=16N\varepsilon=\frac{1}{6N} and TO(ε2(b+dlog(Nd)))T\coloneqq O(\varepsilon^{-2}(b+d\log(Nd))), where d=max{α(Q0+),α(Q0)}d=\max\{\alpha(Q_{0}^{+}),\alpha(Q_{0}^{-})\}, and bb is the maximum bit length of σ\sigma.
4:for t=1,,Tt=1,\dots,T do
5:     \trianglerightLeft normalization
6:     If ΦV(I)Diag(b)trε\lVert\Phi_{V}(I)-\operatorname{Diag}(b^{-})\rVert_{\operatorname{tr}}\leq\varepsilon then return Yes.
7:     Compute the Cholesky decomposition CC=ΦV(I)CC^{\dagger}=\Phi_{V}(I). Set g=Diag(b)1/2C1g=\operatorname{Diag}(b^{-})^{1/2}C^{-1} and update VVg,IV\leftarrow V_{g,I}.
8:     \trianglerightRight normalization
9:     If ΦV(I)Diag(b+)trε\lVert\Phi^{*}_{V}(I)-\operatorname{Diag}(b^{+})\rVert_{\operatorname{tr}}\leq\varepsilon then return Yes.
10:     Compute the Cholesky decomposition CC=ΦV(I)CC^{\dagger}=\Phi_{V}^{*}(I). Set h=Diag(b+)1/2C1h=\operatorname{Diag}(b^{+})^{1/2}C^{-1} and update VVI,hV\leftarrow V_{I,h}.
11:return No.

To run the algorithm, we first need to show how to compute the value of ΦV\Phi_{V} in polynomial time. Note that the naive computation of ΦV\Phi_{V} requires exponential time, as the number of terms in the sum is exponential. However, we can compute ΦV\Phi_{V} in polynomial time using the underlying quiver structure. We show an algorithm in Algorithm 3. The algorithm for the dual map is similar and therefore omitted.

Algorithm 3 Algorithm for computing ΦV(X)\Phi_{V}(X).
1:a representation VV of an acyclic quiver QQ and a block matrix X=(Xs:sQ0+)X=(X_{s}:s\in Q_{0}^{+})
2:Let XiOX_{i}\coloneqq O for iQ0Q0+i\in Q_{0}\setminus Q_{0}^{+}.
3:for aQ1a\in Q_{1} in a topological order do
4:     XhaXha+V(a)XtaV(a)X_{ha}\leftarrow X_{ha}+V(a)X_{ta}V(a)^{\dagger}.
5:return (Xi:iQ0)(X_{i}:i\in Q^{-}_{0}).

Next, we show an upper bound of the accuracy parameter ε\varepsilon that is sufficient to decide the σ\sigma-semistability.

Lemma 3.3.

Let (b+,b)(b^{+},b^{-}) be as above. Let 0<ε16N.0<\varepsilon\leq\frac{1}{6N}. If there exist g,hg,h such that (ΦV)g,h(I)Diag(b)trε\lVert(\Phi_{V})_{g,h}(I)-\operatorname{Diag}(b^{-})\rVert_{\operatorname{tr}}\leq\varepsilon and (ΦV)g,h(I)Diag(b+)trε\lVert(\Phi_{V})_{g,h}^{*}(I)-\operatorname{Diag}(b^{+})\rVert_{\operatorname{tr}}\leq\varepsilon, then VV is σ\sigma-semistable.

Proof.

Let Φ~V\tilde{\Phi}_{V} be the scaled CP map of ΦV\Phi_{V} by g,hg,h. Let b~+,b~\tilde{b}^{+},\tilde{b}^{-} be the spectra of Φ~V(I),Φ~V(I)\tilde{\Phi}_{V}(I),\tilde{\Phi}_{V}^{*}(I), respectively. First, observe that (b~+,b~)(\tilde{b}^{+},\tilde{b}^{-}) is also a feasible marginal. To see this, note that

Φ~V(I)tDiag(b(t))=Ut(Diag(b~)σ(t)NIα(t))Ut,\displaystyle\tilde{\Phi}_{V}(I)_{t}-\operatorname{Diag}(b^{-}(t))=U_{t}\left(\operatorname{Diag}(\tilde{b}^{-})-\frac{\sigma^{-}(t)}{N}I_{\alpha(t)}\right)U_{t}^{\dagger}, (3.13)

where UtU_{t} is the unitary matrix that diagonalizes Φ~V(I)\tilde{\Phi}_{V}(I). Similarly,

Φ~V(I)sDiag(b+(s))=Us(Diag(b~+)σ+(s)NIα(s))Us.\displaystyle\tilde{\Phi}_{V}^{*}(I)_{s}-\operatorname{Diag}(b^{+}(s))=U_{s}\left(\operatorname{Diag}(\tilde{b}^{+})-\frac{\sigma^{+}(s)}{N}I_{\alpha(s)}\right)U_{s}^{\dagger}. (3.14)

Hence, g=(tQ0Ut)1gg^{\prime}={\bigl{(}\bigoplus_{t\in Q_{0}^{-}}U_{t}\bigr{)}}^{-1}g and h=(sQ0+Us)1hh^{\prime}={\bigl{(}\bigoplus_{s\in Q_{0}^{+}}U_{s}\bigr{)}}^{-1}h satisfy the same assumption. By Lemma 2.7, there is no difference between the approximate scalability with upper triangular matrices and nonsingular matrices because (b+,b)(b^{+},b^{-}) is constant within each block. Therefore, (b~+,b~)(\tilde{b}^{+},\tilde{b}^{-}) is a feasible marginal under block triangular scaling. Furthermore,

b~+b+1\displaystyle\big{\lVert}\tilde{b}^{+}-b^{+}\big{\rVert}_{1} Φ~V(I)Diag(b)trε,\displaystyle\leq\big{\lVert}\tilde{\Phi}_{V}(I)-\operatorname{Diag}(b^{-})\big{\rVert}_{\operatorname{tr}}\leq\varepsilon, (3.15)
b~b1\displaystyle\big{\lVert}\tilde{b}^{-}-b^{-}\big{\rVert}_{1} Φ~V(I)Diag(b+)trε.\displaystyle\leq\big{\lVert}\tilde{\Phi}^{*}_{V}(I)-\operatorname{Diag}(b^{+})\big{\rVert}_{\operatorname{tr}}\leq\varepsilon. (3.16)

Since (b~+,b~)(\tilde{b}^{+},\tilde{b}^{-}) is feasible, from Lemma 2.6, we have

tQ0i=1α(t)Δb~(t)idim(LtF(t)i)+sQ0+j=1α(s)Δb~+(s)jdim(RsF+(s)j)B~,\displaystyle\sum_{t\in Q_{0}^{-}}\sum_{i=1}^{\alpha(t)}\Delta\tilde{b}^{-}(t)_{i}\dim(L_{t}\cap F^{-}(t)_{i})+\sum_{s\in Q_{0}^{+}}\sum_{j=1}^{\alpha(s)}\Delta\tilde{b}^{+}(s)_{j}\dim(R_{s}\cap F^{+}(s)_{j})\leq\tilde{B}, (3.17)

where B~\tilde{B} denotes the sum of the entries of b~+\tilde{b}^{+}. Let Δdim(LtF(t)i)dim(LtF(t)i)dim(LtF(t)i1)\Delta\dim(L_{t}\cap F^{-}(t)_{i})\coloneqq\dim(L_{t}\cap F^{-}(t)_{i})-\dim(L_{t}\cap F^{-}(t)_{i-1}), where we conventionally define F(t)0={0}F^{-}(t)_{0}=\{0\}. Similarly, define Δdim(RsF+(s)j)\Delta\dim(R_{s}\cap F^{+}(s)_{j}). Rewriting the LHS, we have

tQ0i=1α(t)b~(t)iΔdim(LtF(t)i)+sQ0+j=1α(s)b~+(s)jΔdim(RsF+(s)j)B~.\displaystyle\sum_{t\in Q_{0}^{-}}\sum_{i=1}^{\alpha(t)}\tilde{b}^{-}(t)_{i}\Delta\dim(L_{t}\cap F^{-}(t)_{i})+\sum_{s\in Q_{0}^{+}}\sum_{j=1}^{\alpha(s)}\tilde{b}^{+}(s)_{j}\Delta\dim(R_{s}\cap F^{+}(s)_{j})\leq\tilde{B}. (3.18)

Observe that

tQ0i=1α(t)b(t)iΔdim(LtF(t)i)\displaystyle\sum_{t\in Q_{0}^{-}}\sum_{i=1}^{\alpha(t)}b^{-}(t)_{i}\Delta\dim(L_{t}\cap F^{-}(t)_{i}) tQ0i=1α(t)(|b(t)ib~(t)i|+b~(t)i)Δdim(LtF(t)i)\displaystyle\leq\sum_{t\in Q_{0}^{-}}\sum_{i=1}^{\alpha(t)}(\lvert b^{-}(t)_{i}-\tilde{b}^{-}(t)_{i}\rvert+\tilde{b}^{-}(t)_{i})\Delta\dim(L_{t}\cap F^{-}(t)_{i}) (3.19)
bb~1+tQ0i=1α(t)b~(t)iΔdim(LtF(t)i)\displaystyle\leq\big{\lVert}b^{-}-\tilde{b}^{-}\big{\rVert}_{1}+\sum_{t\in Q_{0}^{-}}\sum_{i=1}^{\alpha(t)}\tilde{b}^{-}(t)_{i}\Delta\dim(L_{t}\cap F^{-}(t)_{i}) (3.20)
ε+tQ0i=1α(t)b~(t)iΔdim(LtF(t)i).\displaystyle\leq\varepsilon+\sum_{t\in Q_{0}^{-}}\sum_{i=1}^{\alpha(t)}\tilde{b}^{-}(t)_{i}\Delta\dim(L_{t}\cap F^{-}(t)_{i}). (3.21)

where the second inequality follows from Δdim(LtF+(t)i)1\Delta\dim(L_{t}\cap F^{+}(t)_{i})\leq 1. Similarly,

sQ0+j=1α(s)b+(s)jΔdim(RsF+(s)j)ε+sQ0+j=1α(s)b~+(s)jΔdim(RsF+(s)j).\displaystyle\sum_{s\in Q_{0}^{+}}\sum_{j=1}^{\alpha(s)}b^{+}(s)_{j}\Delta\dim(R_{s}\cap F^{+}(s)_{j})\leq\varepsilon+\sum_{s\in Q_{0}^{+}}\sum_{j=1}^{\alpha(s)}\tilde{b}^{+}(s)_{j}\Delta\dim(R_{s}\cap F^{+}(s)_{j}). (3.22)

Furthermore,

B~=sQ0+j=1α(s)b~+(s)jsQ0+j=1α(s)(|b+(s)jb~+(s)j|+b+(s)j)ε+1.\displaystyle\tilde{B}=\sum_{s\in Q_{0}^{+}}\sum_{j=1}^{\alpha(s)}\tilde{b}^{+}(s)_{j}\leq\sum_{s\in Q_{0}^{+}}\sum_{j=1}^{\alpha(s)}(\lvert b^{+}(s)_{j}-\tilde{b}^{+}(s)_{j}\rvert+b^{+}(s)_{j})\leq\varepsilon+1. (3.23)

Combining these inequalities with the assumption that 0<ε1/6N0<\varepsilon\leq 1/6N, we obtain

tQ0i=1α(t)Δb(t)idim(LtF(t)i)+sQ0+j=1α(s)Δb+(s)jdim(RsF+(s)j)1+3ε1+12N.\displaystyle\sum_{t\in Q_{0}^{-}}\sum_{i=1}^{\alpha(t)}\Delta b^{-}(t)_{i}\dim(L_{t}\cap F^{-}(t)_{i})+\sum_{s\in Q_{0}^{+}}\sum_{j=1}^{\alpha(s)}\Delta b^{+}(s)_{j}\dim(R_{s}\cap F^{+}(s)_{j})\leq 1+3\varepsilon\leq 1+\frac{1}{2N}. (3.24)

Therefore, we have

tQ0σ(t)dimLt+sQ0+σ+(s)dimRsN+12.\displaystyle\sum_{t\in Q_{0}^{-}}\sigma^{-}(t)\dim L_{t}+\sum_{s\in Q_{0}^{+}}\sigma^{+}(s)\dim R_{s}\leq N+\frac{1}{2}. (3.25)

By the integrality of the LHS and NN, we can get rid of the 12\frac{1}{2} term. Therefore, we obtain (3.12) and show that VV is σ\sigma-semistable. ∎

Theorem 3.4.

Let VV be a representation of an acyclic quiver QQ with Gaussian integer entries and σ\sigma be a weight with σ(α)=0\sigma(\alpha)=0. Algorithm 2 correctly decides the σ\sigma-semistability of VV in O(N2(b+dlog(Nd)))O(N^{2}(b+d\log(Nd))) iterations, where N=σ+(α)=σ(α)N=\sigma^{+}(\alpha)=\sigma^{-}(\alpha), d=max{α(Q0+),α(Q0)}d=\max\{\alpha(Q_{0}^{+}),\alpha(Q_{0}^{-})\}, and bb is the maximum bit length of σ\sigma. Each iteration can be executed in O(|Q1|αmax3+sQ0+α(s)3+tQ0α(t)3)O(|Q_{1}|\alpha_{\max}^{3}+\sum_{s\in Q_{0}^{+}}\alpha(s)^{3}+\sum_{t\in Q_{0}^{-}}\alpha(t)^{3}) time, where αmax=maxiQ0α(i)\alpha_{\max}=\max_{i\in Q_{0}}\alpha(i).

Proof.

If VV is σ\sigma-semistable, then ΦV\Phi_{V} is approximately scalable to the marginals (b+,b)(b^{+},b^{-}). By Theorem 2.2, the algorithm must find such a scaling within the stated iterations. Consequently, the algorithm outputs Yes. If VV is not σ\sigma-semistable, then there is no scaling g,hg,h such that (ΦV)g,h(I)Diag(b)trε\lVert(\Phi_{V})_{g,h}(I)-\operatorname{Diag}(b^{-})\rVert_{\operatorname{tr}}\leq\varepsilon and (ΦV)g,h(I)Diag(b+)trε\lVert(\Phi_{V})_{g,h}^{*}(I)-\operatorname{Diag}(b^{+})\rVert_{\operatorname{tr}}\leq\varepsilon for ε=1/6N\varepsilon=1/6N by Lemma 3.3. Thus, the algorithm outputs No. This proves the correctness of the algorithm.

The number of iterations is immediate from the algorithm. The time complexity of each iteration is dominated by the computations of ΦV(I)\Phi_{V}(I) and ΦV(I)\Phi_{V}^{*}(I) and that of the block Cholesky decomposition. The former takes O(|Q1|αmax3)O(|Q_{1}|\alpha_{\max}^{3}) time and the latter takes O(tQ0α(t)3+sQ0+α(s)3)O(\sum_{t\in Q_{0}^{-}}\alpha(t)^{3}+\sum_{s\in Q_{0}^{+}}\alpha(s)^{3}) time. ∎

3.3 Finding the extreme maximizer in King’s criterion

In this subsection, we extend the result from the previous section to find the extreme maximizer in King’s criterion. The idea is to use the shrunk subspace algorithm for the linear matrix (3.1).

Let

ΦVσ:iQ0+Mat(α(i))σ+(i)iQ0Mat(α(i))σ(i)\displaystyle\Phi_{V}^{\sigma}:\bigoplus_{i\in Q_{0}^{+}}{\operatorname{Mat}(\alpha(i))}^{\sigma^{+}(i)}\to\bigoplus_{i\in Q_{0}^{-}}{\operatorname{Mat}(\alpha(i))}^{\sigma^{-}(i)} (3.26)

be a CP map that maps X=(Xs,p:sQ0+,p[σ+(s)])X=\bigoplus(X_{s,p}:s\in Q_{0}^{+},p\in[\sigma^{+}(s)]) to

(ΦVσ(X))t,q=sQ0+p[σ+(s)]P:st pathV(P)Xs,pV(P)\displaystyle(\Phi_{V}^{\sigma}(X))_{t,q}=\sum_{s\in Q_{0}^{+}}\sum_{p\in[\sigma^{+}(s)]}\sum_{P:\text{$s$--$t$ path}}V(P)X_{s,p}V(P)^{\dagger} (3.27)

for tQ0t\in Q_{0}^{-} and q[σ(t)]q\in[\sigma^{-}(t)]. Let Jσ+=sQ0+σ+(s)Iα(s)J^{+}_{\sigma}=\bigoplus_{s\in Q_{0}^{+}}\sigma^{+}(s)I_{\alpha(s)} and Jσ=tQ0σ(t)Iα(t)J^{-}_{\sigma}=\bigoplus_{t\in Q_{0}^{-}}\sigma^{-}(t)I_{\alpha(t)}. Note that

(ΦVσ(I))t,q=sQ0+σ+(s)P:st pathV(P)V(P)=ΦV(Jσ+)t,\displaystyle(\Phi_{V}^{\sigma}(I))_{t,q}=\sum_{s\in Q_{0}^{+}}\sigma^{+}(s)\sum_{P:\text{$s$--$t$ path}}V(P)V(P)^{\dagger}=\Phi_{V}(J^{+}_{\sigma})_{t}, (3.28)
((ΦVσ)(I))s,p=tQ0σ(t)P:st pathV(P)V(P)=ΦV(Jσ)s.\displaystyle((\Phi_{V}^{\sigma})^{*}(I))_{s,p}=\sum_{t\in Q_{0}^{-}}\sigma^{-}(t)\sum_{P:\text{$s$--$t$ path}}V(P)^{\dagger}V(P)=\Phi_{V}^{*}(J^{-}_{\sigma})_{s}. (3.29)

Thus, one can compute ΦVσ(I)\Phi_{V}^{\sigma}(I) and (ΦVσ)(I)(\Phi_{V}^{\sigma})^{*}(I) in poly(|Q|,|α|,|σ|)\operatorname{poly}(|Q|,|\alpha|,|\sigma|) time using Algorithm 3. By Lemma 3.1, the minimum shrunk subspace of ΦVσ\Phi_{V}^{\sigma} corresponds to the minimum maximizer in King’s criterion. To find the minimum shrunk subspace of ΦVσ\Phi_{V}^{\sigma}, one can simply use Theorem 2.1, which runs in poly(|Q|,|α|,|σ|,b)\operatorname{poly}(|Q|,|\alpha|,|\sigma|,b) time, because one can compute ΦVσ(I)\Phi_{V}^{\sigma}(I) and (ΦVσ)(I)(\Phi_{V}^{\sigma})^{*}(I) in poly(|Q|,|α|,|σ|,b)\operatorname{poly}(|Q|,|\alpha|,|\sigma|,b) time.

The maximum maximizer can also be found by considering Φ\Phi^{*} instead of Φ\Phi. To see this, observe that the maximum shrunk subspace corresponds to the maximum independent subspace (L,R)(L,R) such that dimL\dim L is the smallest. Since tr(ΠLΦ(ΠR))=tr(Φ(ΠL)ΠR)\operatorname{tr}(\Pi_{L}\Phi(\Pi_{R}))=\operatorname{tr}(\Phi^{*}(\Pi_{L})\Pi_{R}), LL is the minimum shrunk subspace of Φ\Phi^{*}.

Therefore, we obtain the following theorem.

Theorem 3.5.

For a quiver representation VV of an acyclic quiver QQ with the dimension vector α\alpha with Gaussian integer entries and a weight σ\sigma, the minimum and maximum maximizers of King’s criterion can be found in poly(|Q|,|α|,|σ|,b)\operatorname{poly}(|Q|,|\alpha|,|\sigma|,b) time, where bb denotes the bit complexity of VV.

4 Harder-Narasimhan filtration and principal partition of quiver representation

In this section, we introduce the principal partition of a quiver representation based on parametric submodular function minimization and show that it coincides with the HN-filtration.

Let us recall the definition. Let σQ0\sigma\in\mathbb{Z}^{Q_{0}} be a weight and τ+Q0\tau\in\mathbb{Z}_{+}^{Q_{0}} a strictly monotone weight, i.e., a nonnegative weight such that τ(dim¯W)>0\tau(\operatorname{\underline{dim}}W)>0 if W{0}W\neq\{0\}. We define the slope of a quiver nonzero representation VV as μ(V)=σ(dim¯V)/τ(dim¯V)\mu(V)=\sigma(\operatorname{\underline{dim}}V)/\tau(\operatorname{\underline{dim}}V). We say that VV is μ\mu-semistable if μ(W)μ(V)\mu(W)\leq\mu(V) for any nonzero subrepresentation WW of VV. The HN-filtration of VV is a unique filtration {0}=W0<W1<<Wk=V\{0\}=W_{0}<W_{1}<\cdots<W_{k}=V such that (i) μ(Wi/Wi1)>μ(Wi+1/Wi)\mu(W_{i}/W_{i-1})>\mu(W_{i+1}/W_{i}) for i[k1]i\in[k-1] and (ii) Wi/Wi1W_{i}/W_{i-1} is μ\mu-semistable.

4.1 Equivalence to principal partition

Let VV be a quiver representation and λ\lambda\in\mathbb{R} a parameter. Define a parametric modular function fλf_{\lambda} as

fλ(W)λτ(dim¯W)σ(dim¯W).\displaystyle f_{\lambda}(W)\coloneqq\lambda\tau(\operatorname{\underline{dim}}W)-\sigma(\operatorname{\underline{dim}}W). (4.1)

Let (λ)\mathcal{L}(\lambda) denote the modular lattice of the minimizers of fλf_{\lambda}.

Lemma 4.1.

Let λ>λ\lambda>\lambda^{\prime}. If W(λ)W\in\mathcal{L}(\lambda) and W(λ)W^{\prime}\in\mathcal{L}(\lambda^{\prime}), then WWW\leq W^{\prime}.

Proof.

For notational simplicity, we abbreviate σ(dim¯W)\sigma(\operatorname{\underline{dim}}W) and τ(dim¯W)\tau(\operatorname{\underline{dim}}W) as σ(W)\sigma(W) and τ(W)\tau(W), respectively. Using the modularity, we have

fλ(W)+fλ(W)\displaystyle f_{\lambda}(W)+f_{\lambda^{\prime}}(W^{\prime}) =λτ(W)+λτ(W)σ(W)σ(W)\displaystyle=\lambda\tau(W)+\lambda^{\prime}\tau(W^{\prime})-\sigma(W)-\sigma(W^{\prime}) (4.2)
=λ(τ(W)+τ(W))σ(W)σ(W)+(λλ)τ(W)\displaystyle=\lambda^{\prime}(\tau(W)+\tau(W^{\prime}))-\sigma(W)-\sigma(W^{\prime})+(\lambda-\lambda^{\prime})\tau(W) (4.3)
=λ(τ(W+W)+τ(WW))σ(W+W)σ(WW)+(λλ)τ(W)\displaystyle=\lambda^{\prime}(\tau(W+W^{\prime})+\tau(W\cap W^{\prime}))-\sigma(W+W^{\prime})-\sigma(W\cap W^{\prime})+(\lambda-\lambda^{\prime})\tau(W) (4.4)
=fλ(W+W)+fλ(WW)+(λλ)(τ(W)τ(WW))\displaystyle=f_{\lambda^{\prime}}(W+W^{\prime})+f_{\lambda}(W\cap W^{\prime})+(\lambda-\lambda^{\prime})(\tau(W)-\tau(W\cap W^{\prime})) (4.5)
=fλ(W+W)+fλ(WW)+(λλ)τ(W/(WW))\displaystyle=f_{\lambda^{\prime}}(W+W^{\prime})+f_{\lambda}(W\cap W^{\prime})+(\lambda-\lambda^{\prime})\tau(W/(W\cap W^{\prime})) (4.6)
fλ(W)+fλ(W),\displaystyle\geq f_{\lambda^{\prime}}(W^{\prime})+f_{\lambda}(W), (4.7)

where the last inequality follows since τ(W/(WW))0\tau(W/(W\cap W^{\prime}))\geq 0 by the assumptions on τ\tau and W,WW,W^{\prime} are minimizers of fλf_{\lambda} and fλf_{\lambda^{\prime}}, respectively. Therefore, the inequality is tight. Since λ>λ\lambda>\lambda^{\prime} and τ\tau is strictly monotone, W/(WW)W/(W\cap W^{\prime}) must be the zero representation, which implies WWW\leq W^{\prime}. ∎

Remark 4.2.

As shown in the above argument, the positivity of τ\tau is required only for W/(WW)W/(W\cap W^{\prime}) for the minimizers WW and WW^{\prime}. This slightly relaxed condition is necessary for the coarse DM-decomposition in Section 4.3.

Let W(λ)W^{-}(\lambda) and W+(λ)W^{+}(\lambda) be the minimum and maximum minimizers of fλf_{\lambda}, respectively. By the lemma, W+(λ)W(λ)W^{+}(\lambda)\leq W^{-}(\lambda^{\prime}) for λ>λ\lambda>\lambda^{\prime}. There must be a finite set of λ\lambda such that (λ)\mathcal{L}(\lambda) consists of more than one element. We call such a value of λ\lambda a critical value. Let λ1>>λk\lambda_{1}>\cdots>\lambda_{k} be the critical values. Then, they induce a filtration

{0}=W(λ1)<W+(λ1)=W(λ2)<W+(λ2)==W(λk)<W+(λk)=V.\displaystyle\{0\}=W^{-}(\lambda_{1})<W^{+}(\lambda_{1})=W^{-}(\lambda_{2})<W^{+}(\lambda_{2})=\cdots=W^{-}(\lambda_{k})<W^{+}(\lambda_{k})=V. (4.8)

We show that this filtration satisfies the definition of the HN-filtration.

Theorem 4.3.

For each i[k]i\in[k],

  • μ(W+(λi)/W(λi))=λi\mu(W^{+}(\lambda_{i})/W^{-}(\lambda_{i}))=\lambda_{i}, and

  • W+(λi)/W(λi)W^{+}(\lambda_{i})/W^{-}(\lambda_{i}) is μ\mu-semistable.

Thus, the above filtration coincides with the HN-filtration.

Proof.

Since W+(λi)W^{+}(\lambda_{i}) and W(λi)W^{-}(\lambda_{i}) are both minimizers of fλif_{\lambda_{i}}, we have fλi(W+(λi))=fλi(W(λi))f_{\lambda_{i}}(W^{+}(\lambda_{i}))=f_{\lambda_{i}}(W^{-}(\lambda_{i})). Hence

0\displaystyle 0 =λi(τ(W+(λi))τ(W(λi)))σ(W+(λi))σ(W(λi))\displaystyle=\lambda_{i}(\tau(W^{+}(\lambda_{i}))-\tau(W^{-}(\lambda_{i})))-\sigma(W^{+}(\lambda_{i}))-\sigma(W^{-}(\lambda_{i})) (4.9)
=λiτ(W+(λi)/W(λi))σ(W+(λi)/W(λi)).\displaystyle=\lambda_{i}\tau(W^{+}(\lambda_{i})/W^{-}(\lambda_{i}))-\sigma(W^{+}(\lambda_{i})/W^{-}(\lambda_{i})). (4.10)

Thus, μ(W+(λi)/W(λi))=λi\mu(W^{+}(\lambda_{i})/W^{-}(\lambda_{i}))=\lambda_{i}.

For the second item, let WW+(λi)/W(λi)W^{\prime}\leq W^{+}(\lambda_{i})/W^{-}(\lambda_{i}) be a nonzero subrepresentation of W+(λi)/W(λi)W^{+}(\lambda_{i})/W^{-}(\lambda_{i}). Then, W+W(λi)W^{\prime}+W^{-}(\lambda_{i}) is a subrepresentation of VV. Since W(λi)W^{-}(\lambda_{i}) is a minimizer of fλif_{\lambda_{i}}, we have

λiτ(W+W(λi))σ(W+W(λi))λiτ(W(λi))σ(W(λi)).\displaystyle\lambda_{i}\tau(W^{\prime}+W^{-}(\lambda_{i}))-\sigma(W^{\prime}+W^{-}(\lambda_{i}))\geq\lambda_{i}\tau(W^{-}(\lambda_{i}))-\sigma(W^{-}(\lambda_{i})). (4.11)

By the modularity of σ\sigma and τ\tau, we have λiτ(W)σ(W)0,\lambda_{i}\tau(W^{\prime})-\sigma(W^{\prime})\geq 0, which implies μ(W)λi=μ(W+(λi)/W(λi))\mu(W^{\prime})\leq\lambda_{i}=\mu(W^{+}(\lambda_{i})/W^{-}(\lambda_{i})). Thus, W+(λi)/W(λi)W^{+}(\lambda_{i})/W^{-}(\lambda_{i}) is μ\mu-semistable. ∎

4.2 Algorithm for Harder-Narasimhan filtration

The above observation yields an efficient algorithm for finding the HN-filtration. First, we reduce the μ\mu-semistablity checking to σ\sigma-semistability checking.

Lemma 4.4.

Let μ=σ/τ\mu=\sigma/\tau be a slope and VV a quiver representation with μ(V)=p/q\mu(V)=p/q, where pp is an integer and qq is a positive integer. Then, VV is μ\mu-semistable if and only if VV is σ\sigma^{\prime}-semistable for σ=qσpτ\sigma^{\prime}=q\sigma-p\tau.

Proof.

By definition, VV is μ\mu-semistable if and only if

μ(V)τ(W)σ(W)μ(V)τ(V)σ(V)=0\displaystyle\mu(V)\tau(W)-\sigma(W)\geq\mu(V)\tau(V)-\sigma(V)=0 (4.12)

for any subrepresentation WVW\leq V. This is equivalent to

(qσpτ)(W)(qσpτ)(V)=0,\displaystyle(q\sigma-p\tau)(W)\leq(q\sigma-p\tau)(V)=0, (4.13)

which is King’s criterion for the σ\sigma^{\prime}-semistability of VV. ∎

Let

S{p/q:p[σ(α),σ+(α)],q[τ(α)]}\displaystyle S\coloneqq\{p/q:p\in[-\sigma^{-}(\alpha),\sigma^{+}(\alpha)],\ q\in[\tau(\alpha)]\} (4.14)

be the set of possible slope values. Note that |S|=O(|σ||τ|)\lvert S\rvert=O(|\sigma||\tau|). For each slope λ=p/q\lambda=p/q, one can compute the maximum minimizer of fλf_{\lambda} by Theorem 3.5 for σ\sigma^{\prime}-semistability. Thus, we obtain the HN-filtration by O(|σ||τ|)O(|\sigma||\tau|) computations of the maximizers of King’s criterion. A pseudocode is given in Algorithm 4.

Algorithm 4 Algorithm for finding the HN-filtration.
1:Set i0i\leftarrow 0, W0{0}W_{0}\leftarrow\{0\}.
2:for each possible slope λ=p/qS\lambda=p/q\in S in the decreasing order do
3:     Invoke the algorithm of Theorem 3.5 with weight σ=qσpτ\sigma^{\prime}=q\sigma-p\tau to find the maximum minimizer WW of fλf_{\lambda}.
4:     if Wi<WW_{i}<W then
5:         Wi+1WW_{i+1}\leftarrow W and ii+1i\leftarrow i+1.      
6:return (W0,,Wi)(W_{0},\dots,W_{i}).
Theorem 4.5.

Let VV be a quiver representation of an acyclic quiver QQ with the dimension vector α\alpha with Gaussian integer entries. Let μ=σ/τ\mu=\sigma/\tau be a slope. Then, Algorithm 4 finds the HN-filtration in poly(|Q|,|α|,|σ|,|τ|,b)\operatorname{poly}(|Q|,|\alpha|,|\sigma|,|\tau|,b) time, where bb is the bit complexity of VV.

4.3 Relation to coarse DM-decomposition

The DM-decomposition is the decomposition of a bipartite graph into smaller graphs that represent the structure of all maximum matchings and minimum vertex covers. It is well known that the DM-decomposition corresponds to a maximal chain of minimizers of the surplus function. Also, this is the finest decomposition of generic matrices under permutation of rows and columns. See [Mur09] for the details.

Generalizing the DM-decomposition, [HS24] introduced the coarse DM-decomposition for linear matrices. Let A=kxkAkA=\sum_{k}x_{k}A_{k} be a linear matrix and 𝒜\mathcal{A} the corresponding matrix space. Without loss of generality, we can assume that AkA_{k} have no common element in their kernels; otherwise, we can simply delete the zero columns of AA corresponding to the common kernel space.

Consider a polytope in 2\mathbb{R}^{2} defined by

conv{(dimX,dimY):(X,Y) independent subspace of A}.\displaystyle\operatorname{conv}\{(\dim X,\dim Y):\text{$(X,Y)$ independent subspace of $A$}\}. (4.15)

Then, the extreme points of the polytope other than the origin correspond to a flag in the lattice of the independent subspace:

n\displaystyle\mathbb{C}^{n} =X0>X1>>X={0},\displaystyle=X_{0}>X_{1}>\dots>X_{\ell}=\{0\}, (4.16)
{0}\displaystyle\{0\} =Y0<Y1<<Y=n.\displaystyle=Y_{0}<Y_{1}<\dots<Y_{\ell}=\mathbb{C}^{n}. (4.17)

By elementary transformation, one can assume that both XiX_{i} and YiY_{i} are coordinate subspaces of the columns and rows, respectively. The coarse DM-decomposition is given by the decomposition of the rows and columns of AA into (Xi1Xi,YiYi1)(X_{i-1}\setminus X_{i},Y_{i}\setminus Y_{i-1}) for i=1,,i=1,\dots,\ell.

Here, we show that the coarse DM-decomposition is a HN-filtration for the generalized Kronecker quiver. Let us regard AA as a representation VV of the generalized Kronecker quiver. Consider weights σ=(1,0)\sigma=(1,0) and τ=(0,1)\tau=(0,1). The HN-filtration is given by the minimizers of a parameterized modular function

fλ(W)=λdimW(2)dimW(1),\displaystyle f_{\lambda}(W)=\lambda\dim W(2)-\dim W(1), (4.18)

where λ\lambda\in\mathbb{R} and WVW\leq V. If λ<0\lambda<0, then the unique minimizer of fλf_{\lambda} is (n,n)(\mathbb{C}^{n},\mathbb{C}^{n}). Suppose that λ0\lambda\geq 0. Since 𝒜W(1)W(2)\mathcal{A}W(1)\leq W(2) for any subrepresentation WW, we can assume that W(2)=𝒜W(1)W(2)=\mathcal{A}W(1) for any minimizer of fλf_{\lambda}. By the assumption that AA has no common kernel element, W(2){0}W(2)\neq\{0\} if W(1){0}W(1)\neq\{0\}. So τ\tau is strictly monotone for minimizers of fλf_{\lambda} (see Remark 4.2). Therefore, it suffices to consider a parameterized submodular function

fλ(U)=λdim𝒜UdimU,\displaystyle f_{\lambda}(U)=\lambda\dim\mathcal{A}U-\dim U, (4.19)

where λ0\lambda\geq 0 and UnU\leq\mathbb{C}^{n} is a subspace. Let λ1>>λk0\lambda_{1}>\cdots>\lambda_{k}\geq 0 be the critical values. As before, denote by UiU_{i}^{-} and Ui+U_{i}^{+} the minimum and maximum minimizers of fλif_{\lambda_{i}}, respectively. They form a flag in the column space, as follows:

{0}=U1<U1+=U2<U2+==Uk<Uk+=n.\displaystyle\{0\}=U_{1}^{-}<U_{1}^{+}=U_{2}^{-}<U_{2}^{+}=\cdots=U_{k}^{-}<U_{k}^{+}=\mathbb{C}^{n}. (4.20)

Rename the flag as n=Y0>Y1>>Yk={0}\mathbb{C}^{n}=Y_{0}>Y_{1}>\dots>Y_{k}=\{0\} to match the notation of the coarse DM-decomposition. Let Xi=𝒜(Yi)X_{i}=\mathcal{A}(Y_{i})^{\perp} (i=0,1,,ki=0,1,\dots,k) be a flag in the row space, where ZZ^{\perp} denotes the orthogonal complement of a vector space ZZ. Then, (Xi,Yi)(X_{i},Y_{i}) is an independent subspace of AA by construction. In fact, the above flags are exactly the flags used in the coarse DM-decomposition. To see this, let us consider an independent subspace (X,Y)(X,Y) that corresponds to an extreme point of the polytope. As (dimX,dimY)(\dim X,\dim Y) is extreme, there exists a slope (λ,1)(\lambda,1) that exposes it. That is,

λdimX+dimYλdimX+dimY\displaystyle\lambda\dim X+\dim Y\geq\lambda\dim X^{\prime}+\dim Y^{\prime} (4.21)

for any independent (X,Y)(X^{\prime},Y^{\prime}). Substituting X=𝒜(Y)X=\mathcal{A}(Y)^{\perp},

λ(ndim𝒜(Y))+dimYλ(ndim𝒜(Y))+dimY,\displaystyle\lambda(n-\dim\mathcal{A}(Y))+\dim Y\geq\lambda(n-\dim\mathcal{A}(Y^{\prime}))+\dim Y^{\prime}, (4.22)

which shows that YY minimizes fλf_{\lambda} and vice versa.

Therefore, Algorithm 4 finds the coarse DM-decomposition of AA. Since |τ|=|σ|=1\lvert\tau\rvert=\lvert\sigma\rvert=1, Algorithm 4 runs in polynomial time.

Remark 4.6.

[HS24] defined the coarse DM-decomposition for any field. Since the underlying quiver is the generalized Kronecker quiver, finding the maximizers of King’s criterion is exactly finding the minimum shrunk subspaces of an approximate matrix space. Therefore, instead of modified operator Sinkhorn iteration, one can use an algebraic algorithm of [IQS18] which can work in any field.

5 Rank-one representation

In this section, we particularly focus on a rank-one representation VV of an acyclic quiver Q=(Q0,Q1)Q=(Q_{0},Q_{1}), i.e., for each arc aQ1a\in Q_{1}, the corresponding linear map V(a)V(a) is rank-one. For a weight σQ0\sigma\in\mathbb{Z}^{Q_{0}}, we present a simple combinatorial characterization of the σ\sigma-semistability of VV in terms of linear matroids and show that the σ\sigma-semistability of VV is equivalent to the feasibility of a certain instance of the submodular flow problem arising from (V,σ)(V,\sigma) (with easily verifiable additional conditions). In particular, these imply that one can check the σ\sigma-semistability of a rank-one representation in strongly polynomial time.

5.1 Preliminarlies on matroid and submodular flow

We first recall the basics of matroids. A matroid (e.g., [Oxl11]) is a pair 𝐌=(S,)\mathbf{M}=(S,\mathcal{B}) of a finite set SS and a nonempty family 2S\mathcal{B}\subseteq 2^{S} of subsets satisfying the following exchange axiom: for any B,BB,B^{\prime}\in\mathcal{B} and xBBx\in B\setminus B^{\prime}, there exists xBBx^{\prime}\in B^{\prime}\setminus B such that B{x}{x}B\setminus\{x\}\cup\{x^{\prime}\}\in\mathcal{B}. The family \mathcal{B} is referred to as the base family of 𝐌\mathbf{M} and each member in \mathcal{B} a base. For a matroid 𝐌=(S,)\mathbf{M}=(S,\mathcal{B}), its rank function r:2Sr:2^{S}\to\mathbb{Z} is defined by r(X)max{BX:B}r(X)\coloneqq\max\{B\cap X:B\in\mathcal{B}\}. It is well-known that the rank function is submodular, i.e., r(X)+r(Y)r(XY)+r(XY)r(X)+r(Y)\geq r(X\cap Y)+r(X\cup Y) for X,YSX,Y\subseteq S.

Example 5.1 (linear matroid).

One of the most fundamental examples of matroids is a linear matroid, which arises from a matrix, or equivalently, a finite multiset of vectors, as follows. Let UU be a vector space and SUS\subseteq U a finite multiset of vectors. Then, the pair (S,)(S,\mathcal{B}) in which

{XS:|X|=dimX=dimS}\displaystyle\mathcal{B}\coloneqq\{X\subseteq S:|X|=\dim\langle X\rangle=\dim\langle S\rangle\} (5.1)

forms a matroid; we refer to this as a linear matroid generated by SS. The rank function rr of a linear matroid is simply the map XdimXX\mapsto\dim\langle X\rangle.

We then prepare notation before introducing the submodular flow problem. For a submodular function f:2S{+}f:2^{S}\to\mathbb{R}\cup\{+\infty\} with f()=0f(\emptyset)=0 and f(S)<+f(S)<+\infty, its base polyhedron 𝐁(f)\mathbf{B}(f) is a polyhedron defined by

𝐁(f){xS:x(S)=f(S),x(X)f(X)(XS)}.\displaystyle\mathbf{B}(f)\coloneqq\{x\in\mathbb{R}^{S}:x(S)=f(S),\ x(X)\leq f(X)\ (\forall X\subseteq S)\}. (5.2)

The following are fundamental to the theory of submodular functions:

Lemma 5.2 ([Mur03, Chapter 4]).

Let f:2S1{+}f:2^{S_{1}}\to\mathbb{R}\cup\{+\infty\} and g:2S2{+}g:2^{S_{2}}\to\mathbb{R}\cup\{+\infty\} be submodular functions satisfying that f()=g()=0f(\emptyset)=g(\emptyset)=0 and f(S1)f(S_{1}) and g(S2)g(S_{2}) are finite.

  1. (1)

    If ff is integer-valued, then 𝐁(f)\mathbf{B}(f) is an integral polyhedron.

  2. (2)

    If S1=S2S_{1}=S_{2}, then 𝐁(f+g)=𝐁(f)+𝐁(g)\mathbf{B}(f+g)=\mathbf{B}(f)+\mathbf{B}(g), where “++” in the right-hand side denotes the Minkowski sum. If S1S_{1} and S2S_{2} are disjoint, then 𝐁(f+g)=𝐁(f)×𝐁(g)\mathbf{B}(f+g)=\mathbf{B}(f)\times\mathbf{B}(g).

Example 5.3 (base polytope of a matroid rank function).

Let rr be the rank function of a matroid 𝐌=(S,)\mathbf{M}=(S,\mathcal{B}) and (r)#(-r)^{\#} the dual [Fuj05, Section 2.3] of r-r, i.e.,

(r)#(X)r(SX)r(S);\displaystyle(-r)^{\#}(X)\coloneqq r(S\setminus X)-r(S); (5.3)

both rr and (r)#(-r)^{\#} are submodular. Then, the base polytopes 𝐁(r)\mathbf{B}(r) and 𝐁((r)#)\mathbf{B}((-r)^{\#}) are represented as

𝐁(r)=conv{χB:B},𝐁((r)#)=conv{χB:B},\displaystyle\mathbf{B}(r)=\operatorname{conv}\left\{\chi_{B}:B\in\mathcal{B}\right\},\qquad\mathbf{B}((-r)^{\#})=\operatorname{conv}\left\{-\chi_{B}:B\in\mathcal{B}\right\}, (5.4)

where χB{0,1}S\chi_{B}\in\{0,1\}^{S} denotes the characteristic vector of BSB\subseteq S, i.e., χB(s)=1\chi_{B}(s)=1 if sSs\in S and χB(s)=0\chi_{B}(s)=0 if sSBs\in S\setminus B.

In addition, let k+k\in\mathbb{Z}_{+} be a nonnegative integer. Then, by Lemma 5.2 (2), we have

𝐁(kr)\displaystyle\mathbf{B}(kr) =conv{=1kχB:B1,,Bk},\displaystyle=\operatorname{conv}\left\{\sum_{\ell=1}^{k}\chi_{B_{\ell}}:B_{1},\dots,B_{k}\in\mathcal{B}\right\}, (5.5)
𝐁((kr)#)\displaystyle\mathbf{B}((-kr)^{\#}) =conv{=1kχB:B1,,Bk}.\displaystyle=\operatorname{conv}\left\{-\sum_{\ell=1}^{k}\chi_{B_{\ell}}:B_{1},\dots,B_{k}\in\mathcal{B}\right\}. (5.6)

Moreover, by Lemma 5.2 (1), the set of integral points in 𝐁(kr)\mathbf{B}(kr) (resp. 𝐁((kr)#)\mathbf{B}((-kr)^{\#})) consists of i=1kχBi\sum_{i=1}^{k}\chi_{B_{i}} (resp. i=1kχBi-\sum_{i=1}^{k}\chi_{B_{i}}) for B1,,BkB_{1},\dots,B_{k}\in\mathcal{B}, that is,

𝐁(kr)S\displaystyle\mathbf{B}(kr)\cap\mathbb{Z}^{S} ={=1kχB:B1,,Bk},\displaystyle=\left\{\sum_{\ell=1}^{k}\chi_{B_{\ell}}:B_{1},\dots,B_{k}\in\mathcal{B}\right\}, (5.7)
𝐁((kr)#)S\displaystyle\mathbf{B}((-kr)^{\#})\cap\mathbb{Z}^{S} ={=1kχB:B1,,Bk}.\displaystyle=\left\{-\sum_{\ell=1}^{k}\chi_{B_{\ell}}:B_{1},\dots,B_{k}\in\mathcal{B}\right\}. (5.8)

In (the feasibility version of) the submodular flow problem [EG77] (see also [Fuj05, Section 5.1]), given a directed graph D=(S,A)D=(S,A), an upper capacity function c¯:A{+}\overline{c}:A\to\mathbb{R}\cup\{+\infty\}, a lower capacity function c¯:A{}\underline{c}:A\to\mathbb{R}\cup\{-\infty\}, and a submodular function f:2S{+}f:2^{S}\to\mathbb{R}\cup\{+\infty\} with f()=f(S)=0f(\emptyset)=f(S)=0, we are asked to find a feasible submodular flow, namely, a flow (vector) φA\varphi\in\mathbb{R}^{A} such that c¯(a)φ(a)c¯(a)\underline{c}(a)\leq\varphi(a)\leq\overline{c}(a) for each aAa\in A and its boundary φ\partial\varphi belongs to the base polyhedron 𝐁(f)\mathbf{B}(f), if it exists. We say that an instance (D,c¯,c¯,f)(D,\overline{c},\underline{c},f) is feasible if there is a feasible submodular flow for (D,c¯,c¯,f)(D,\overline{c},\underline{c},f).

The following is well-known in combinatorial optimization.

Theorem 5.4 ([Fra84]; see also [Fuj05, Section 5.3]).

Let D=(S,A)D=(S,A) be a directed graph, c¯:A{+}\overline{c}:A\to\mathbb{R}\cup\{+\infty\} an upper capacity, c¯:A{}\underline{c}:A\to\mathbb{R}\cup\{-\infty\} a lower capacity, and f:2S{+}f:2^{S}\to\mathbb{R}\cup\{+\infty\} a submodular function with f()=f(S)=0f(\emptyset)=f(S)=0. Then, (D,c¯,c¯,f)(D,\overline{c},\underline{c},f) is feasible if and only if

c¯(Out(X))c¯(In(X))+f(VX)0\displaystyle\overline{c}(\operatorname{Out}(X))-\underline{c}(\operatorname{In}(X))+f(V\setminus X)\geq 0 (5.9)

holds for all XSX\subseteq S. In particular, if ff is integer-valued, then there exists an integral feasible submodular flow φA\varphi\in\mathbb{Z}^{A} whenever it is feasible.

5.2 Reduction from semistability to submodular flow

Let Q=(Q0,Q1)Q=(Q_{0},Q_{1}) be an acyclic quiver and VV a rank-one representation of QQ. Then for each arc aQ1a\in Q_{1}, the corresponding linear map V(a)V(a) is representable as vafav_{a}f_{a} for some nonzero vector vaV(ha)v_{a}\in V(ha) and nonzero dual vector faV(ta)f_{a}\in V(ta)^{*}, i.e., V(a)V(a) is the map V(ta)uvafa(u)V(ha)V(ta)\ni u\mapsto v_{a}f_{a}(u)\in V(ha). Building on this representation, we consider that the data of VV consist of

  • a vector space V(i)V(i) for each vertex iQ0i\in Q_{0},

  • a nonzero vector vaV(i)v_{a}\in V(i) for each vertex iQ0i\in Q_{0} and incoming arc aIn(i)a\in\operatorname{In}(i), and

  • a nonzero dual vector faV(i)f_{a}\in V(i)^{*} for each vertex iQ0i\in Q_{0} and outgoing arc aOut(i)a\in\operatorname{Out}(i).

Let Si+{fa:aOut(i)}S_{i}^{+}\coloneqq\{f_{a}:a\in\operatorname{Out}(i)\} and Si{va:aIn(i)}S_{i}^{-}\coloneqq\{v_{a}:a\in\operatorname{In}(i)\} for iQ0i\in Q_{0}. Note that Si+S_{i}^{+} and SiS_{i}^{-} may be multisets; even if faf_{a} and fbf_{b} are the same vectors for distinct a,bOut(i)a,b\in\operatorname{Out}(i), we distinguish faf_{a} from fbf_{b} in Si+S_{i}^{+}.

We can easily observe the following from the definition of a subrepresentation.

Lemma 5.5.

A tuple (W(i))iQ0(W(i))_{i\in Q_{0}} of vector subspaces W(i)V(i)W(i)\leq V(i) induces a subrepresentation WW of VV if and only if for any arc aQ1a\in Q_{1} with W(ta)kerfaW(ta)\not\leq\ker f_{a}, we have vaW(ha)\langle v_{a}\rangle\leq W(ha).

The above observation (Lemma 5.5) gives rise to the directed graph D[V]=(S,A)D[V]=(S,A) defined by

SiQ0(Si+Si),\displaystyle S\coloneqq\bigcup_{i\in Q_{0}}\left(S_{i}^{+}\cup S_{i}^{-}\right), (5.10)
A{(fa,va):aQ1}{(va,fb):iQ0,vaSi,fbSi+,vakerfb}.\displaystyle A\coloneqq\left\{(f_{a},v_{a}):a\in Q_{1}\right\}\cup\left\{(v_{a},f_{b}):i\in Q_{0},\ v_{a}\in S_{i}^{-},\ f_{b}\in S_{i}^{+},\ v_{a}\notin\ker f_{b}\right\}. (5.11)

Intuitively, an arc in AA of the form (fa,va)(f_{a},v_{a}) represents the original arc aQ1a\in Q_{1} of the quiver, whereas that of the form (va,fb)(v_{a},f_{b}) represents “if W(ha)=W(tb)W(ha)=W(tb) contains vav_{a} then W(tb)kerfbW(tb)\not\leq\ker f_{b}; hence, W(hb)W(hb) must contain vbv_{b} for WW to be a subrepresentation of VV.” See Figure 2.

Refer to caption
Figure 2: The left is an original quiver QQ and the right is the directed graph D[V]D[V] constructed from QQ and a rank-one representation VV of QQ. The red circles and blue squares in the right graph represent the vertices of D[V]D[V] corresponding to the outgoing arcs (endowed with nonzero dual vector) and incoming arcs (endowed with nonzero vector) in QQ, respectively.

This digraph D[V]D[V] enables us to characterize the σ\sigma-semistability of VV in a combinatorial manner. For each iQ0i\in Q_{0}, let 𝐌i+=(Si+,i+)\mathbf{M}_{i}^{+}=(S_{i}^{+},\mathcal{B}_{i}^{+}) (resp. 𝐌i=(Si,i)\mathbf{M}_{i}^{-}=(S_{i}^{-},\mathcal{B}_{i}^{-})) denote the linear matroid generated by Si+S_{i}^{+} (resp. SiS_{i}^{-}). Moreover, let ri+r_{i}^{+} (resp. rir_{i}^{-}) denote the rank function of 𝐌i+\mathbf{M}_{i}^{+} (resp. 𝐌i\mathbf{M}_{i}^{-}). Then, we obtain the following.

Theorem 5.6.

Let VV be a rank-one representation of QQ and σQ0\sigma\in\mathbb{Z}^{Q_{0}} a weight. Then VV is σ\sigma-semistable if and only if

  1. (K1)

    iQ0σ+(i)dimV(i)=iQ0σ(i)dimV(i)Σ\sum_{i\in Q_{0}}\sigma^{+}(i)\dim V(i)=\sum_{i\in Q_{0}}\sigma^{-}(i)\dim V(i)\eqqcolon\Sigma and

  2. (K2)

    for any lower set XSX\subseteq S of D[V]D[V], we have

    iQ0(σ+(i)ri+(Si+X)+σ(i)ri(SiX))Σ.\displaystyle\sum_{i\in Q_{0}}\left(\sigma^{+}(i)r_{i}^{+}(S_{i}^{+}\setminus X)+\sigma^{-}(i)r_{i}^{-}(S_{i}^{-}\cap X)\right)\geq\Sigma. (5.12)
Proof.

By King’s criterion, VV is σ\sigma-semistable if and only if σ(dim¯V)=0\sigma(\operatorname{\underline{dim}}V)=0 and σ(dim¯W)0\sigma(\operatorname{\underline{dim}}W)\leq 0 for any subrepresentation WW of VV. The former condition σ(dim¯V)=0\sigma(\operatorname{\underline{dim}}V)=0 of King’s criterion is equivalent to (K1). Hence, it suffices to see the equivalence between the latter condition of King’s criterion and (K2) under the condition (K1).

Suppose that the latter condition of King’s criterion holds, i.e., σ(dim¯W)0\sigma(\operatorname{\underline{dim}}W)\leq 0 for any subrepresentation WW of VV. Take any lower set XSX\subseteq S of D[V]D[V]. We define from XX, the tuple (W(i))iQ0(W(i))_{i\in Q_{0}} of vector subspaces W(i)V(i)W(i)\leq V(i) by

W(i){{kerfb:fbSi+X}if σ(i)0,va:vaSiXif σ(i)<0.\displaystyle W(i)\coloneqq\begin{cases}\bigcap\{\ker f_{b}:f_{b}\in S_{i}^{+}\setminus X\}&\text{if $\sigma(i)\geq 0$},\\ \langle v_{a}:v_{a}\in S_{i}^{-}\cap X\rangle&\text{if $\sigma(i)<0$}.\end{cases} (5.13)

We prove that (W(i))iQ0(W(i))_{i\in Q_{0}} induces a subrepresentation of VV. By Lemma 5.5, it suffices to see that for any arc aQ1a\in Q_{1} with vaW(ha)\langle v_{a}\rangle\not\leq W(ha), or equivalently, vaW(ha)v_{a}\notin W(ha), we have W(ta)kerfaW(ta)\leq\ker f_{a}. Since XX is a lower set, there is no arc between vaSiXv_{a}\in S_{i}^{-}\cap X and fbSi+Xf_{b}\in S_{i}^{+}\setminus X, i.e., every vaSiXv_{a}\in S_{i}^{-}\cap X is in kerfb\ker f_{b}. Hence, for any iQ0i\in Q_{0}, we have va:vaSiX{kerfb:fbSi+X}\langle v_{a}:v_{a}\in S_{i}^{-}\cap X\rangle\leq\bigcap\{\ker f_{b}:f_{b}\in S_{i}^{+}\setminus X\}, which implies va:vaSiXW(i)\langle v_{a}:v_{a}\in S_{i}^{-}\cap X\rangle\leq W(i). Take any arc aQ1a\in Q_{1} with vaW(ha)v_{a}\notin W(ha). Then, vaShaXv_{a}\in S_{ha}^{-}\setminus X by the above argument, and hence faSta+Xf_{a}\in S_{ta}^{+}\setminus X as XX is a lower set. Therefore, W(ta)kerfaW(ta)\leq\ker f_{a} holds.

Since (W(i))iQ0(W(i))_{i\in Q_{0}} induces a subrepresentation of VV, we have σ(dim¯W)0\sigma(\operatorname{\underline{dim}}W)\leq 0, i.e.,

iQ0(σ+(i)dim{kerfb:fbSi+X}σ(i)dimva:vaSiX)0.\displaystyle\sum_{i\in Q_{0}}\left(\sigma^{+}(i)\dim\bigcap\{\ker f_{b}:f_{b}\in S_{i}^{+}\setminus X\}-\sigma^{-}(i)\dim\langle v_{a}:v_{a}\in S_{i}^{-}\cap X\rangle\right)\leq 0. (5.14)

Since dim{kerfb:fbSi+X}=dimV(i)dimfb:fbSi+X=dimV(i)ri+(Si+X)\dim\bigcap\{\ker f_{b}:f_{b}\in S_{i}^{+}\setminus X\}=\dim V(i)^{*}-\dim\langle f_{b}:f_{b}\in S_{i}^{+}\setminus X\rangle=\dim V(i)-r_{i}^{+}(S_{i}^{+}\setminus X) and dimva:vaSiX=ri(SiX)\dim\langle v_{a}:v_{a}\in S_{i}^{-}\cap X\rangle=r_{i}^{-}(S_{i}^{-}\cap X) (see Example 5.1 for the rank function of a linear matroid), the inequality (5.14) is equivalent to

iQ0(σ+(i)ri+(Si+X)+σ(i)ri(SiX))Σ,\displaystyle\sum_{i\in Q_{0}}\left(\sigma^{+}(i)r_{i}^{+}(S_{i}^{+}\setminus X)+\sigma^{-}(i)r_{i}^{-}(S_{i}^{-}\cap X)\right)\geq\Sigma, (5.15)

where we recall that Σ=iQ0σ+(i)dimV(i)\Sigma=\sum_{i\in Q_{0}}\sigma^{+}(i)\dim V(i) by (K1).

Conversely, suppose that (K2) holds. Take any subrepresentation WW of VV. Our aim is to prove σ(dim¯W)0\sigma(\operatorname{\underline{dim}}W)\leq 0.

We define a vertex subset XSX\subseteq S of D[V]D[V] by

XiQ0({vaSi:vaW(i)}{fbSi+:W(i)kerfb}).\displaystyle X\coloneqq\bigcup_{i\in Q_{0}}\left(\{v_{a}\in S_{i}^{-}:v_{a}\in W(i)\}\cup\{f_{b}\in S_{i}^{+}:W(i)\not\leq\ker f_{b}\}\right). (5.16)

Then, XX forms a lower set of D[V]D[V]. Indeed, if fbStb+Xf_{b}\in S_{tb}^{+}\cap X, or equivalently, W(tb)kerfbW(tb)\not\leq\ker f_{b}, then W(hb)W(hb) contains vbv_{b} since WW is a subrepresentation of VV, implying vbShbXv_{b}\in S_{hb}^{-}\cap X. If vaShaXv_{a}\in S_{ha}^{-}\cap X, that is, vaW(ha)v_{a}\in W(ha), then W(ha)kerfbW(ha)\not\leq\ker f_{b} for (va,fb)A(v_{a},f_{b})\in A by the definition of arc (va,fb)(v_{a},f_{b}); thus we obtain fbStb+X(=Sha+X)f_{b}\in S_{tb}^{+}\cap X(=S_{ha}^{+}\cap X) for any (va,fb)A(v_{a},f_{b})\in A.

By reversing the argument from (5.14) to (5.15) above, we obtain the inequality (5.14) for this XX. Since va:vaSiXW(i){kerfb:fbSi+X}\langle v_{a}:v_{a}\in S_{i}^{-}\cap X\rangle\leq W(i)\leq\bigcap\{\ker f_{b}:f_{b}\in S_{i}^{+}\setminus X\} and σ+(i),σ(i)0\sigma^{+}(i),\sigma^{-}(i)\geq 0 for iQ0i\in Q_{0}, the LHS of (5.14) is at least σ(dim¯W)\sigma(\operatorname{\underline{dim}}W). Therefore, we conclude that σ(dim¯W)0\sigma(\operatorname{\underline{dim}}W)\leq 0. ∎

Theorem 5.6 implies the following necessary conditions for σ\sigma-semistability.

Corollary 5.7.

Let VV be a rank-one representation of QQ and σQ0\sigma\in\mathbb{Z}^{Q_{0}} a weight. If VV is σ\sigma-semistable, then (K1) and the following full-dimensional condition (F) hold:

  1. (F)

    ri+(Si+)=dimV(i)r_{i}^{+}(S_{i}^{+})=\dim V(i) for iQ0+i\in Q_{0}^{+} and ri(Si)=dimV(i)r_{i}^{-}(S_{i}^{-})=\dim V(i) for iQ0i\in Q_{0}^{-}.

Proof.

We clearly have ri+(Si+)dimV(i)=dimV(i)r_{i}^{+}(S_{i}^{+})\leq\dim V(i)^{*}=\dim V(i) and ri(Si)dimV(i)r_{i}^{-}(S_{i}^{-})\leq\dim V(i) for each iQ0i\in Q_{0}. Hence we obtain iQ0σ+(i)ri+(Si+)Σ\sum_{i\in Q_{0}}\sigma^{+}(i)r_{i}^{+}(S_{i}^{+})\leq\Sigma and iQ0σ(i)ri(Si)Σ\sum_{i\in Q_{0}}\sigma^{-}(i)r_{i}^{-}(S_{i}^{-})\leq\Sigma.

Suppose that VV is σ\sigma-semistable. Then, by Theorem 5.6, we have (K1) and iQ0σ+(i)ri+(Si+)Σ\sum_{i\in Q_{0}}\sigma^{+}(i)r_{i}^{+}(S_{i}^{+})\geq\Sigma (corresponding to the lower set X=X=\emptyset in (K2)) and iQ0σ(i)ri(Si)Σ\sum_{i\in Q_{0}}\sigma^{-}(i)r_{i}^{-}(S_{i}^{-})\geq\Sigma (corresponding to the lower set X=SX=S in (K2)). Thus we obtain iQ0σ+(i)ri+(Si+)=iQ0σ(i)ri(Si)=Σ\sum_{i\in Q_{0}}\sigma^{+}(i)r_{i}^{+}(S_{i}^{+})=\sum_{i\in Q_{0}}\sigma^{-}(i)r_{i}^{-}(S_{i}^{-})=\Sigma, which implies ri+(Si+)=dimV(i)r_{i}^{+}(S_{i}^{+})=\dim V(i) for iQ0+i\in Q_{0}^{+} and ri(Si)=dimV(i)r_{i}^{-}(S_{i}^{-})=\dim V(i) for iQ0i\in Q_{0}^{-}. ∎

In the following, we see that the combinatorial characterization of the σ\sigma-semistability given in Theorem 5.6 can be rephrased as the feasibility of a certain instance of the submodular flow problem on D[V]D[V]. Here, we assume that the conditions (K1) and (F) hold (recall Corollary 5.7). Set c¯\overline{c} and c¯\underline{c} as c¯(a)+\overline{c}(a)\coloneqq+\infty and c¯(a)0\underline{c}(a)\coloneqq 0 for aAa\in A, denoted as ¯\overline{\infty} and 0¯\underline{0}, respectively. We define a function fV:2Sf_{V}:2^{S}\to\mathbb{R} by

fV(X)\displaystyle f_{V}(X)\coloneqq iQ0(σ+(i)ri+(Si+X)+(σ(i)ri)#(SiX))\displaystyle\sum_{i\in Q_{0}}\left(\sigma^{+}(i)r_{i}^{+}(S_{i}^{+}\cap X)+(-\sigma^{-}(i)r_{i}^{-})^{\#}(S_{i}^{-}\cap X)\right) (5.17)
=\displaystyle{}= iQ0(σ+(i)ri+(Si+X)+(σ(i)ri(SiX)σ(i)ri(Si)))\displaystyle\sum_{i\in Q_{0}}\left(\sigma^{+}(i)r_{i}^{+}(S_{i}^{+}\cap X)+\left(\sigma^{-}(i)r_{i}^{-}(S_{i}^{-}\setminus X)-\sigma^{-}(i)r_{i}^{-}(S_{i}^{-})\right)\right) (5.18)
=\displaystyle{}= iQ0(σ+(i)ri+(Si+X)+σ(i)ri(SiX))Σ\displaystyle\sum_{i\in Q_{0}}\left(\sigma^{+}(i)r_{i}^{+}(S_{i}^{+}\cap X)+\sigma^{-}(i)r_{i}^{-}(S_{i}^{-}\setminus X)\right)-\Sigma (5.19)

for XSX\subseteq S, where the second equality follows from the definition of the dual (σ(i)ri)#(-\sigma^{-}(i)r_{i}^{-})^{\#} (see Example 5.3) and the last follows from iQ0σ(i)ri(Si)=Σ\sum_{i\in Q_{0}}\sigma^{-}(i)r_{i}^{-}(S_{i}^{-})=\Sigma by (K1) and (F). The assumptions (K1) and (F) also imply f()=f(S)=0f(\emptyset)=f(S)=0. Since the rank functions ri+r_{i}^{+} and rir_{i}^{-} are submodular and σ+(i)\sigma^{+}(i) and σ(i)\sigma^{-}(i) are nonnegative, the function fVf_{V} is also submodular. Thus (D[V],¯,0¯,fV)(D[V],\overline{\infty},\underline{0},f_{V}) forms an instance of the submodular flow problem.

The following theorem says that the σ\sigma-semistability can be characterized as the feasibility of (D[V],¯,0¯,fV)(D[V],\overline{\infty},\underline{0},f_{V}). Here, for a finite set TT, subset TTT^{\prime}\subseteq T, and γT\gamma\in\mathbb{R}^{T}, let γ|T\gamma|_{T^{\prime}} denote the projection of γ\gamma to T\mathbb{R}^{T^{\prime}}.

Theorem 5.8.

Let VV be a rank-one representation of QQ and σQ0\sigma\in\mathbb{Z}^{Q_{0}} a weight. Then, VV is σ\sigma-semistable if and only if it satisfies (K1), (F), and

  1. (S)

    the instance (D[V],¯,0¯,fV)(D[V],\overline{\infty},\underline{0},f_{V}) of the submodular flow problem is feasible.

In addition, the last condition (S) can be replaced by the following more combinatorial condition (S)’:

  1. (S)’

    there is a nonnegative integral flow φ+A\varphi\in\mathbb{Z}_{+}^{A} such that, for each iQ0i\in Q_{0}, we have φ|Si+==1σ+(i)χB\partial\varphi|_{S_{i}^{+}}=\sum_{\ell=1}^{\sigma^{+}(i)}\chi_{B_{\ell}} for some B1,,Bσ+(i)i+B_{1},\dots,B_{\sigma^{+}(i)}\in\mathcal{B}_{i}^{+} and φ|Si==1σ(i)χB\partial\varphi|_{S_{i}^{-}}=-\sum_{\ell=1}^{\sigma^{-}(i)}\chi_{B_{\ell}} for some B1,,Bσ(i)iB_{1},\dots,B_{\sigma^{-}(i)}\in\mathcal{B}_{i}^{-}.

Proof.

We first see that, under (K1) and (F), the condition (K2) is equivalent to (S), which implies the former assertion. By Theorem 5.4, the rank-one representation VV satisfies (S) if and only if

¯(Out(X))0¯(In(X))+fV(VX)0\displaystyle\overline{\infty}(\operatorname{Out}(X))-\underline{0}(\operatorname{In}(X))+f_{V}(V\setminus X)\geq 0 (5.20)

for any XSX\subseteq S, which can be rephrased using (5.19) as

iQ0(σ+(i)ri+(Si+(i)X)+σ(i)ri(SiX))Σ\displaystyle\sum_{i\in Q_{0}}\left(\sigma^{+}(i)r_{i}^{+}(S_{i}^{+}(i)\setminus X)+\sigma^{-}(i)r_{i}^{-}(S_{i}^{-}\cap X)\right)\geq\Sigma (5.21)

for any lower set XSX\subseteq S of D[V]D[V]. This is equivalent to the condition (K2).

It follows from Lemma 5.2 (2), Example 5.3, and the fact that fVf_{V} is integer-valued that

𝐁(fV)S=iQ0{=1σ+(i)χB:B1,,Bσ+(i)i+}×{=1σ(i)χB:B1,,Bσ(i)i}.\displaystyle\mathbf{B}(f_{V})\cap\mathbb{Z}^{S}=\prod_{i\in Q_{0}}\left\{\sum_{\ell=1}^{\sigma^{+}(i)}\chi_{B_{\ell}}:B_{1},\dots,B_{\sigma^{+}(i)}\in\mathcal{B}_{i}^{+}\right\}\times\left\{-\sum_{\ell=1}^{\sigma^{-}(i)}\chi_{B_{\ell}}:B_{1},\dots,B_{\sigma^{-}(i)}\in\mathcal{B}_{i}^{-}\right\}. (5.22)

From the above and Theorem 5.4, we obtain the latter assertion. ∎

The conditions (K1) and (F) are clearly verifiable in strongly polynomial time. Since we can minimize the submodular function XiQ0(σ+(i)ri+(Si+X)+σ(i)ri(SiX))X\mapsto\sum_{i\in Q_{0}}\left(\sigma^{+}(i)r_{i}^{+}(S_{i}^{+}\setminus X)+\sigma^{-}(i)r_{i}^{-}(S_{i}^{-}\cap X)\right) over the ring family {X:X is a lower set of D[V]}\{X:\text{$X$ is a lower set of $D[V]$}\} in strongly polynomial time (see [Jia21] and the references therein), the condition (K2) is also verifiable in strongly polynomial time. Similarly, one can check (S) in strongly polynomial time, since the function Xc¯(Out(X))c¯(In(X))+f(VX)X\mapsto\overline{c}(\operatorname{Out}(X))-\underline{c}(\operatorname{In}(X))+f(V\setminus X) in Theorem 5.4 is submodular [Fuj05, Section 2.3]; see also [Fra84]. Therefore, we obtain the following:

Theorem 5.9.

Let VV be a rank-one representation of a quiver QQ and σQ0\sigma\in\mathbb{Z}^{Q_{0}} a weight. Then, one can check if VV is σ\sigma-semistable in strongly polynomial time.

5.3 Implications

Theorem 5.8 states that King’s criterion for a rank-one representation serves as a good characterization for the existence of a feasible submodular flow in a certain instance of the submodular flow problem, or equivalently, a flow such that its boundary can be decomposed as a sum of indicators of matroid bases. Pursuing this direction, we specializes Theorem 5.8 to quivers having specific structures: generalized Kronecker quivers and star quivers, which arise from linear matroid intersection and rank-one BL polytopes, respectively. In particular, we observe that Theorem 5.8 can recover well-known good characterizations on these problems.

Generalized Kronecker quivers and linear matroid intersction.

Suppose that Q=(Q0,Q1)Q=(Q_{0},Q_{1}) is a generalized Kronecker quiver, i.e., Q0Q_{0} consists of the two vertices 1,21,2 and Q1Q_{1} consists of mm parallel arcs a1,,ama_{1},\dots,a_{m} from 11 to 22 (recall Figure 1). Let VV be a rank-one representation of QQ, in which V(ak)vkfkV(a_{k})\coloneqq v_{k}f_{k} for some nonzero vkV(2)v_{k}\in V(2) and fkV(1)f_{k}\in V(1)^{*}. Here, fkf_{k} is regarded as a row vector. That is, S+S1+={f1,,fm}S^{+}\coloneqq S_{1}^{+}=\{f_{1},\dots,f_{m}\} and SS2={v1,,vm}S^{-}\coloneqq S_{2}^{-}=\{v_{1},\dotsc,v_{m}\} (these may be multisets). Let 𝐌+𝐌1+\mathbf{M}^{+}\coloneqq\mathbf{M}_{1}^{+} (resp. 𝐌𝐌2\mathbf{M}^{-}\coloneqq\mathbf{M}_{2}^{-}) be the linear matroid generated by S+S^{+} (resp. SS^{-}), and r+r1+r^{+}\coloneqq r_{1}^{+} (resp. rr2r^{-}\coloneqq r_{2}^{-}) denotes the rank function of 𝐌+\mathbf{M}^{+} (resp. 𝐌\mathbf{M}^{-}). We assume that S+=V(1)\langle S^{+}\rangle=V(1)^{*} and S=V(2)\langle S^{-}\rangle=V(2). In the following, we naturally identify S+S^{+} and SS^{-} with Q1Q_{1} via the correspondences between fi,vif_{i},v_{i} and aia_{i}, that is, we consider that both of the ground sets of 𝐌+\mathbf{M}^{+} and 𝐌\mathbf{M}^{-} are Q1Q_{1}.

Set a weight σ\sigma as σ=(1,1)\sigma=(1,-1). Then, King’s criterion is that dimV(1)=dimV(2)\dim V(1)=\dim V(2) and dimUdim(k=1mV(ak)U)\dim U\leq\dim(\sum_{k=1}^{m}V(a_{k})U) for any subspace UV(1)U\leq V(1) (see Example 1.2). A feasible integral submodular flow in this setting corresponds to bases of 𝐌+\mathbf{M}^{+} and 𝐌\mathbf{M}^{-} indexed by the same arc sets. Hence, we obtain the following corollary of Theorem 5.8, which agrees with the well-known characterization of the existence of a common base in linear matroid intersection according to Lovász [Lov89].

Corollary 5.10 ([Lov89]).

Assume dimV(1)=dimV(2)\dim V(1)=\dim V(2). Then, there is a common base BQ1B\subseteq Q_{1} of 𝐌+\mathbf{M}^{+} and 𝐌\mathbf{M}^{-} if and only if dimUdim(k=1mV(k)U)\dim U\leq\dim(\sum_{k=1}^{m}V(k)U) for any subspace UV(1)U\leq V(1).

Note that the latter condition in Corollary 5.10 is equivalent to the nc-nonsingularity of a linear symbolic matrix k=1mxkvkfk\sum_{k=1}^{m}x_{k}v_{k}f_{k}, as described in Example 1.2.

Star quivers and rank-one BL polytopes.

Suppose that Q=(Q0,Q1)Q=(Q_{0},Q_{1}) is a star quiver, i.e., Q0={0,1,,m}Q_{0}=\{0,1,\dots,m\} and Q1={(0,1),,(0,m)}Q_{1}=\{(0,1),\dots,(0,m)\} (recall Figure 1). Let VV be a rank-one representation of QQ, in which V((0,i))vifiV((0,i))\coloneqq v_{i}f_{i} for some nonzero viV(i)v_{i}\in V(i) and fiV(0)f_{i}\in V(0)^{*}. That is, S0+={f1,,fm}S_{0}^{+}=\{f_{1},\dots,f_{m}\} (this may be a multiset) and Si={vi}S_{i}^{-}=\{v_{i}\} for each i[m]i\in[m]. Recall that 𝐌𝐌0+\mathbf{M}\coloneqq\mathbf{M}_{0}^{+} is the linear matroid generated by S0+S_{0}^{+} and rr0+r\coloneqq r_{0}^{+} denotes the rank function of 𝐌\mathbf{M}. We assume S0+=V(0)\langle S_{0}^{+}\rangle=V(0)^{*} and dimV(i)=1\dim V(i)=1 for i[m]i\in[m].

Let σ=(d,c1,,cm)Q0\sigma=(d,-c_{1},\dotsc,-c_{m})\in\mathbb{Z}^{Q_{0}} be a weight with d,c1,,cm>0d,c_{1},\dotsc,c_{m}>0. Then, as described in Example 1.3, King’s criterion is that (c1/d,,cm/d)(c_{1}/d,\dots,c_{m}/d) is in the Brascamp–Lieb (BL) polytope determined from the rank-one BL-datum (f1,,fm)(f_{1},\dots,f_{m}), which is the set of points p+mp\in\mathbb{R}_{+}^{m} such that

dimUi=1mpidim(fiU)\displaystyle\dim U\leq\sum_{i=1}^{m}p_{i}\dim(f_{i}U) (5.23)

for all subspaces UV(0)U\leq V(0). On the other hand, feasible submodular flows correspond to the base polytope of 𝐌\mathbf{M}. Therefore, we obtain the following corollary of Theorem 5.8, which recovers the characterization of the rank-one BL polytope by [Bar98].

Corollary 5.11 ([Bar98]).

The BL polytope associated with the rank-one BL-datum (f1,,fm)(f_{1},\dots,f_{m}) is the base polytope of 𝐌\mathbf{M}.

6 Polynomial-time semistability testing for general quivers

In this section, we present a polynomial-time algorithm for checking the semistability of a quiver representation of a general quiver, possibly having directed cycles, under the GL-action.666As mentioned in Section 1.5, after submitting the first version of this paper, we were informed by an anonymous reviewer that a similar approach for general quivers has been sketched in [Mul17, Theorem 10.8 and the last paragraph of Section 10.2 ].

6.1 Algebraic condition for semistability

Let Q=(Q0,Q1)Q=(Q_{0},Q_{1}) be a general quiver and VV a representation of QQ. The semistability of general quivers can be defined in the same way as that of acyclic quivers, which were given in Section 1.1. Our starting point for checking the semistability is the following Le Bruyn–Procesi theorem. Note that a path means a directed walk, i.e., a path can visit each vertex many times, as mentioned in Section 1.

Theorem 6.1 ([BP90]).

The invariant ring of the action of GL(Q,α)\operatorname{GL}(Q,\alpha) on the representation space of a quiver QQ is generated by

tr[V(ak)V(a2)V(a1)]\displaystyle\operatorname{tr}[V(a_{k})\cdots V(a_{2})V(a_{1})] (6.1)

for a closed path (a1,a2,,ak)(a_{1},a_{2},\dots,a_{k}) of length k1k\geq 1 in QQ. Furthermore, closed paths with length 1kα(Q0)21\leq k\leq\alpha(Q_{0})^{2} generate the invariant ring, where α=dim¯V\alpha=\operatorname{\underline{dim}}V.

By Theorem 6.1, to check if VV is semistable, it suffices to check if trV(C)0\operatorname{tr}V(C)\neq 0 for some closed path CC with a maximum length of α(Q0)2{\alpha(Q_{0})}^{2} in QQ, where V(C)V(ak)V(a2)V(a1)V(C)\coloneqq V(a_{k})\dotsb V(a_{2})V(a_{1}) if C=(a1,a2,,ak)C=(a_{1},a_{2},\dots,a_{k}).

Recall that a cycle in a digraph can be detected by repeatedly multiplying the adjacency matrix. Let AA be the adjacency matrix of QQ. It is well-known that QQ has an iijj path of length kk if and only if (Ak)ij0(A^{k})_{ij}\neq 0. Thus, QQ has a cycle of length kk if and only if (Ak)ii0(A^{k})_{ii}\neq 0 for some vertex ii.

One can generalize this to quiver representations. Define the adjacency matrix of a quiver representation VV as a partitioned matrix

A:iQ0α(i)iQ0α(i)\displaystyle A:\bigoplus_{i\in Q_{0}}\mathbb{C}^{\alpha(i)}\to\bigoplus_{i\in Q_{0}}\mathbb{C}^{\alpha(i)} (6.2)

whose (i,j)(i,j)-block is given by

aQ0:ta=j,ha=ixaV(a),\displaystyle\sum_{a\in Q_{0}:ta=j,ha=i}x_{a}V(a), (6.3)

where xax_{a} is an indeterminate that is pairwise noncommutative with other indeterminates but commutes with complex numbers.

Lemma 6.2.

Let VV be a representation of a quiver QQ and AA the adjacency matrix of VV. Then, VV is semistable if and only if

k=1α(Q0)2trAk\displaystyle\sum_{k=1}^{{\alpha(Q_{0})}^{2}}\operatorname{tr}A^{k} (6.4)

is a nonzero polynomial, where α=dim¯V\alpha=\operatorname{\underline{dim}}V.

Proof.

Clearly, for iQ0i\in Q_{0} and k[α(Q0)2]k\in[{\alpha(Q_{0})}^{2}], the (i,i)(i,i)-block of AkA^{k} equals

C: closed path of length k starting at ixCV(C),\displaystyle\sum_{\text{$C$: closed path of length $k$ starting at $i$}}x^{C}V(C), (6.5)

where xC=xakxa1x^{C}=x_{a_{k}}\dotsb x_{a_{1}} if C=(a1,,ak)C=(a_{1},\dots,a_{k}). Thus, taking the trace within the (i,i)(i,i)-block, we obtain a polynomial

C: closed path of length k starting at ixCtrV(C),\displaystyle\sum_{\text{$C$: closed path of length $k$ starting at $i$}}x^{C}\operatorname{tr}V(C), (6.6)

which is nonzero if and only if there is a closed path CC of length kk starting at ii such that trV(C)0\operatorname{tr}V(C)\neq 0. Thus, by Theorem 6.1, VV is semistable if and only if the trace of the (i,i)(i,i)-block of AkA^{k} is nonzero for some iQ0i\in Q_{0} and k[α(Q0)2]k\in[{\alpha(Q_{0})}^{2}]. It is easily checked that the polynomial (6.4) is the sum of such traces. Furthermore, xCx^{C} are different monic monomials for different closed paths CC, and the leftmost factor in xCx^{C} corresponds to the first arc that appears in CC. Here, the sum of the distinct traces does not cancel out, proving the claim. ∎

Therefore, to check the semistability of VV, it suffices to check if (6.4) is a noncommutative polynomial. Section 6.2 provides a deterministic algorithm for this by using an identity testing algorithm for noncommutative algebraic branching programs.

6.2 Deterministic algorithm via white-box polynomial identity testing

A (noncommutative) algebraic branching program (ABP) consists of a directed acyclic graph whose vertices are partitioned into d+1d+1 parts L0,,LdL_{0},\dotsc,L_{d}, each of which is called a layer. The first and last layers L0L_{0} and LdL_{d} consist of a singleton, and the unique vertices in L0L_{0} and LdL_{d} are called the source and sink, respectively. Arcs may only go from LkL_{k} to Lk+1L_{k+1} for k=0,,d1k=0,\dotsc,d-1. Each arc aa is labeled with a homogeneous linear form in noncommutative variables x1,,xnx_{1},\dotsc,x_{n}. The polynomial computed at a vertex vv is the sum over all paths, from the source to vv, of the product of the labeled homogeneous linear forms. An ABP is said to compute a polynomial ff if ff is computed at the sink. The size of an ABP means the number of vertices. Raz and Shpilka [RS05] showed the following result on the identity testing for noncommutative ABPs.

Theorem 6.3 ([RS05, Theorem 4]).

There is an algorithm that, given a noncommutative ABP of size ss in nn indeterminates, verifies whether the ABP computes a zero polynomial in time O(s5+sn4)O(s^{5}+sn^{4}).

The following lemma constructs a polynomial-sized ABP that computes (6.4).

Lemma 6.4.

There is an ABP of size O(α(Q0)4)O({\alpha(Q_{0})}^{4}) that computes the polynomial (6.4).

Proof.

We construct α(Q0)2+3{\alpha(Q_{0})}^{2}+3 layers {s}=L1,L0,,Lα(Q0)2,Lα(Q0)2+1={t}\{s\}=L_{-1},L_{0},\dotsc,L_{{\alpha(Q_{0})}^{2}},L_{{\alpha(Q_{0})}^{2}+1}=\{t\} as follows. Let ss and tt be the source and sink, respectively. Every intermediate layer LkL_{k} (k=0,,α(Q0)2)(k=0,\dotsc,{\alpha(Q_{0})}^{2}) consists of α(Q0)2{\alpha(Q_{0})}^{2} vertices vk,p,qv_{k,p,q} (p,q[α(Q0)])(p,q\in[{\alpha(Q_{0})}]). The source ss connects to v0,p,pv_{0,p,p} for p[α(Q0)]p\in[\alpha(Q_{0})] with the label 11. For k=1,,α(Q0)2k=1,\dotsc,{\alpha(Q_{0})}^{2} and p,q[α(Q0)]p,q\in[{\alpha(Q_{0})}], the incoming arcs to vl,p,qv_{l,p,q} are those from vk1,p,rv_{k-1,p,r} with labels ArqA_{rq} for r[α(Q0)]r\in[{\alpha(Q_{0})}], where the row (column) set of AA is identified with [α(Q0)][\alpha(Q_{0})]. We can inductively check that the polynomial computed at vk,p,qv_{k,p,q} is the (p,q)(p,q) entry in AkA^{k}. We further add an extra vertex vkv^{*}_{k} to LkL_{k} for k=2,,α(Q0)2k=2,\dotsc,{\alpha(Q_{0})}^{2} and connect from vk1,p,pv_{k-1,p,p} to vkv^{*}_{k} with label 11 for every p[α(Q0)]p\in[\alpha(Q_{0})]. We also draw arcs from vα(Q0)2,p,pv_{{\alpha(Q_{0})}^{2},p,p} to tt for all pp in the same way. Then, vkv^{*}_{k} computes trAk1\operatorname{tr}A^{k-1} and tt computes trAα(Q0)2\operatorname{tr}A^{{\alpha(Q_{0})}^{2}}. These polynomials can be aggregated to tt by appending arcs from vkv^{*}_{k} to vk+1v^{*}_{k+1} for k=2,,α(Q0)21k=2,\dotsc,{\alpha(Q_{0})}^{2}-1 with label 1 and from vα(Q0)2v^{*}_{{\alpha(Q_{0})}^{2}} to tt with label 1. This completes the proof. ∎

By Theorem 6.3 and Lemma 6.4, we obtain the following.

Theorem 6.5.

We can check the semistability of a representation VV of a quiver Q=(Q0,Q1)Q=(Q_{0},Q_{1}) in O(α(Q0)20+α(Q0)2ω2|Q1|)O({\alpha(Q_{0})}^{20}+{\alpha(Q_{0})}^{2\omega-2}|Q_{1}|) time, where α=dim¯V\alpha=\operatorname{\underline{dim}}V and ω\omega denotes the exponential of the complexity of matrix multiplication.

Proof.

Let AijOut(i)In(j)A_{ij}\coloneqq\operatorname{Out}(i)\cap\operatorname{In}(j) be the set of arcs from iQ0i\in Q_{0} to jQ0j\in Q_{0}. To reduce the number of arcs, we first find BijAijB_{ij}\subseteq A_{ij} such that {V(a):aBij}\{V(a):a\in B_{ij}\} is a base of {V(a):aAij}\langle\{V(a):a\in A_{ij}\}\rangle for each i,jQ0i,j\in Q_{0}. Let Q=(Q0,Q1)Q^{\prime}=(Q_{0},Q^{\prime}_{1}) with Q1i,jQ0BijQ^{\prime}_{1}\coloneqq\cup_{i,j\in Q_{0}}B_{ij} and VV^{\prime} a representation of QQ^{\prime} naturally obtained from VV by restricting the arc set to Q1Q^{\prime}_{1}. We show that VV is semistable on QQ if and only if VV^{\prime} is on QQ^{\prime}. By Theorem 6.1, it suffices to show that trV(C)0\operatorname{tr}V(C)\neq 0 for some closed path CC of QQ if and only if trV(C)=trV(C)0\operatorname{tr}V^{\prime}(C^{\prime})=\operatorname{tr}V(C^{\prime})\neq 0 for some closed path CC^{\prime} of QQ^{\prime}. The “if” part is clear as CC^{\prime} is also a closed path of QQ. To show the “only if” part, suppose that trV(C)=0\operatorname{tr}V(C^{\prime})=0 for all closed paths CC^{\prime} of QQ^{\prime}. This means that, for any sequence i0,i1,,ik1,ik=i0Q0i_{0},i_{1},\dotsc,i_{k-1},i_{k}=i_{0}\in Q_{0} of vertices,

tr[(aBi0i1xaV(a))(aBi1i2xaV(a))(aBik1i0xaV(a))]\displaystyle\operatorname{tr}\left[\left(\sum_{a\in B_{i_{0}i_{1}}}x_{a}V(a)\right)\left(\sum_{a\in B_{i_{1}i_{2}}}x_{a}V(a)\right)\dotsm\left(\sum_{a\in B_{i_{k-1}i_{0}}}x_{a}V(a)\right)\right] (6.7)

is a zero polynomial. Therefore, replacing the llth factor in (6.7) with any linear combination of V(a)V(a) over aBil1ila\in B_{i_{l-1}i_{l}} for each l[k]l\in[k] cannot make the trace nonzero; hence, trV(C)=0\operatorname{tr}V(C)=0 holds for all closed paths CC of QQ.

Through Gaussian elimination with fast matrix multiplication, the quiver QQ^{\prime} can be obtained in O(|Aij|(α(i)α(j))ω1)=O(|Aij|α(Q0)2ω2)O(|A_{ij}|{(\alpha(i)\alpha(j))}^{\omega-1})=O(|A_{ij}|{\alpha(Q_{0})}^{2\omega-2}) time for each i,ji,j and O(|Q1|α(Q0)2ω2)O(|Q_{1}|{\alpha(Q_{0})}^{2\omega-2}) time in total. We then construct the ABP for QQ^{\prime} promised by Lemma 6.4 and apply Theorem 6.3 to the zero testing. Since the size of the ABP is O(α(Q0)4)O({\alpha(Q_{0})}^{4}) and the number of variables, which is the number of arcs in QQ^{\prime}, is at most i,jQ0α(i)α(j)=α(Q0)2\sum_{i,j\in Q_{0}}\alpha(i)\alpha(j)={\alpha(Q_{0})}^{2}, the running time of the zero testing is O(α(Q0)20)O({\alpha(Q_{0})}^{20}). ∎

Acknowledgments

The authors thank Hiroshi Hirai and Keiya Sakabe for their valuable comments on an earlier version of this paper. The last author thanks Cole Franks for bringing the submodularity of quiver representations to his attention. The authors thank an anonymous reviewer for pointing out the reference [Mul17]. This work was supported by JSPS KAKENHI Grant Number JP24K21315, Japan. The first author was supported by JSPS KAKENHI Grant Numbers JP22K17854, JP24K02901, Japan. The second author was supported by JST, ERATO Grant Number JPMJER1903, JST, CREST Grant Number JPMJCR24Q2, JST, FOREST Grant Number JPMJFR232L, and JSPS KAKENHI Grant Number JP22K17853, Japan. The last author was supported by JPSP KAKENHI Grant Number JP19K20212, and JST, PRESTO Grant Number JPMJPR24K5, Japan.

References

  • [AKRS21] Carlos Améndola, Kathlén Kohn, Philipp Reichenbach and Anna Seigal “Invariant Theory and Scaling Algorithms for Maximum Likelihood Estimation” In SIAM Journal on Applied Algebra and Geometry 5.2, 2021, pp. 304–337 DOI: 10.1137/20M1328932
  • [Bar98] Franck Barthe “On a Reverse Form of the Brascamp-Lieb Inequality” In Inventiones Mathematicae 134.2, 1998, pp. 335–361 DOI: 10.1007/s002220050267
  • [BCCT08] Jonathan Bennett, Anthony Carbery, Michael Christ and Terence Tao “The Brascamp-Lieb Inequalities: Finiteness, Structure and Extremals” In Geometric and Functional Analysis 17.5, 2008, pp. 1343–1415 DOI: 10.1007/s00039-007-0619-6
  • [BP90] Lieven Le Bruyn and Claudio Procesi “Semisimple Representations of Quivers” In Transactions of the American Mathematical Society 317.2, 1990, pp. 585 DOI: 10.2307/2001477
  • [BFGO+18] Peter Burgisser, Cole Franks, Ankit Garg, Rafael Oliveira, Michael Walter and Avi Wigderson “Efficient Algorithms for Tensor Scaling, Quantum Marginals, and Moment Polytopes” In Proceedings of the IEEE 59th Annual Symposium on Foundations of Computer Science (FOCS), 2018, pp. 883–897 DOI: 10.1109/FOCS.2018.00088
  • [BFGO+19] Peter Bürgisser, Cole Franks, Ankit Garg, Rafael Oliveira, Michael Walter and Avi Wigderson “Towards a Theory of Non-Commutative Optimization: Geodesic 1st and 2nd Order Methods for Moment Maps and Polytopes” In Proceedings of the 60th IEEE Annual Symposium on Foundations of Computer Science (FOCS), 2019, pp. 845–861 DOI: 10.1109/FOCS.2019.00055
  • [Che24] Chi-Yu Cheng “A Deterministic Algorithm for Harder–Narasimhan Filtrations for Representations of Acyclic Quivers” In Algebra & Number Theory 18.2, 2024, pp. 319–347 DOI: 10.2140/ant.2024.18.319
  • [CCK22] Calin Chindris, Brett Collins and Daniel Kline “Membership in moment cones and quiver semi-invariants for bipartite quivers” In arXiv, 2022
  • [CD21] Calin Chindris and Harm Derksen “The Capacity of Quiver Representations and Brascamp-Lieb Constants” In International Mathematics Research Notices 2022.24, 2021, pp. 19399–19430 DOI: 10.1093/imrn/rnab064
  • [CI22] Calin Chindris and Jasim Ismaeel “A Quiver Invariant Theoretic Approach to Radial Isotropy and the Paulsen Problem for Matrix Frames” In SIAM Journal on Applied Algebra and Geometry 6.4, 2022, pp. 536–562 DOI: 10.1137/21M141470X
  • [CK21] Calin Chindris and Daniel Kline “Simultaneous robust subspace recovery and semi-stability of quiver representations” In Journal of Algebra 577, 2021, pp. 210–236 DOI: 10.1016/j.jalgebra.2021.03.005
  • [Coh95] P.. Cohn “Skew Fields: Theory of General Division Rings”, Encyclopedia of Mathematics and its Applications Cambridge University Press, 1995
  • [DM17] Harm Derksen and Visu Makam “Polynomial Degree Bounds for Matrix Semi-Invariants” In Advances in Mathematics 310, 2017, pp. 44–63 DOI: 10.1016/j.aim.2017.01.018
  • [DW17] Harm Derksen and Jerzy Weyman “An Introduction to Quiver Representations” American Mathematical Soc., 2017
  • [EG77] Jack Edmonds and Rick Giles “A Min-Max Relation for Submodular Functions on Graphs” In Annals of Discrete Mathematics 1 Elsevier, 1977, pp. 185–204
  • [FR04] Marc Fortin and Christophe Reutenauer “Commutative/noncommutative rank of linear matrices and subspaces of matrices of low rank.” In Séminaire Lotharingien de Combinatoire [electronic only] 52, 2004, pp. B52f–12
  • [Fra84] András Frank “Finding feasible vectors of Edmonds-Giles polyhedra” In Journal of Combinatorial Theory, Series B 36.3 Elsevier BV, 1984, pp. 221–239
  • [Fra18] Cole Franks “Operator Scaling with Specified Marginals” In Proceedings of the 50th Annual ACM SIGACT Symposium on Theory of Computing (STOC), 2018, pp. 190–203 DOI: 10.1145/3188745.3188932
  • [FM23] Cole Franks and Visu Makam “iPCA and stability of star quivers” In arXiv, 2023 DOI: 10.48550/arXiv.2302.09658
  • [FSG23] Cole Franks, Tasuku Soma and Michel X Goemans “Shrunk subspaces via operator Sinkhorn iteration” In Proceedings of the 2023 Annual ACM-SIAM Symposium on Discrete Algorithms (SODA), 2023, pp. 1655–1668
  • [FM20] William Cole Franks and Ankur Moitra “Rigorous Guarantees for Tyler’s M-Estimator via Quantum Expansion” In Proceedings of the 33rd Conference on Learning Theory (COLT) 125 PMLR, 2020, pp. 1601–1632
  • [Fuj05] Satoru Fujishige “Submodular Functions and Optimization” Amsterdam: Elsevier, 2005
  • [Fuj09] Satoru Fujishige “Theory of Principal Partitions Revisited” In Research Trends in Combinatorial Optimization: Bonn 2008 Springer Berlin Heidelberg, 2009, pp. 127–162 DOI: 10.1007/978-3-540-76796-1_7
  • [Gal57] David Gale “A theorem on flows in networks.” In Pacific Journal of Mathematics 7.2, 1957, pp. 1073–1082
  • [GGOW18] Ankit Garg, Leonid Gurvits, Rafael Oliveira and Avi Wigderson “Algorithmic and Optimization Aspects of Brascamp-Lieb Inequalities, via Operator Scaling” In Geometric and Functional Analysis 28.1, 2018, pp. 100–145 DOI: 10.1007/s00039-018-0434-2
  • [GGOW19] Ankit Garg, Leonid Gurvits, Rafael Oliveira and Avi Wigderson “Operator Scaling: Theory and Applications” In Foundations of Computational Mathematics, 2019 DOI: 10.1007/s10208-019-09417-z
  • [Gur04] Leonid Gurvits “Classical Complexity and Quantum Entanglement” In Journal of Computer and System Sciences 69.3, 2004, pp. 448–484 DOI: 10.1016/j.jcss.2004.06.003
  • [HH21] Masaki Hamada and Hiroshi Hirai “Computing the Nc-Rank via Discrete Convex Optimization on CAT(0) Spaces” In SIAM Journal on Applied Algebra and Geometry 5.3, 2021, pp. 455–478 DOI: 10.1137/20M138836X
  • [HN75] G Harder and M S Narasimhan “On the cohomology groups of moduli spaces of vector bundles on curves” In Mathematische Annalen 212.3 Springer Nature, 1975, pp. 215–248
  • [HA02] Lutz Hille and José Antonio de la Peña “Stable representations of quivers” In Journal of Pure and Applied Algebra 172.2, 2002, pp. 205–224 DOI: https://doi.org/10.1016/S0022-4049(01)00167-0
  • [HS24] Hiroshi Hirai and Keiya Sakabe “Gradient Descent for Unbounded Convex Functions on Hadamard Manifolds and its Applications to Scaling Problems” In Proceedings of the IEEE 65th Annual Symposium on Foundations of Computer Science (FOCS), 2024, pp. 2387–2402 DOI: 10.1109/FOCS61266.2024.00139
  • [Hus21] Alana Huszar “Non-Commutative Rank and Semi-Stability of Quiver Representations” In arXiv, 2021 DOI: 10.48550/arXiv.2111.00039
  • [IQS18] Gábor Ivanyos, Youming Qiao and K V Subrahmanyam “Constructive Non-Commutative Rank Computation Is in Deterministic Polynomial Time” In Computational Complexity 27.4, 2018, pp. 561–593 DOI: 10.1007/s00037-018-0165-7
  • [Jia21] Haotian Jiang “Minimizing convex functions with integral minimizers” In Proceedings of the 32nd Annual ACM-SIAM Symposium on Discrete Algorithms (SODA) Society for IndustrialApplied Mathematics, 2021, pp. 976–985
  • [Kin94] Alastair D King “Moduli of representations of finite dimensional algebras” In The Quarterly Journal of Mathematics 45.4 Citeseer, 1994, pp. 515–530
  • [KV18] Bernhard Korte and Jens Vygen “Combinatorial Optimization: Theory and Algorithms” Springer-Verlag, Berlin, 2018 DOI: 10.1007/978-3-662-56039-6
  • [Lov89] László Lovász “Singular Spaces of Matrices and Their Application in Combinatorics” In Bulletin of the Brazilian Mathematical Society 20, 1989, pp. 87–99
  • [Mul17] Ketan Mulmuley “Geometric complexity theory V: Efficient algorithms for Noether normalization” In Journal of the American Mathematical Society 30.1, 2017, pp. 225–309
  • [Mur03] Kazuo Murota “Discrete Convex Analysis” Philadelphia: SIAM, 2003
  • [Mur09] Kazuo Murota “Matrices and Matroids for System Analysis” Springer-Verlag, Berlin, 2009
  • [Oxl11] James Oxley “Matroid Theory” Oxford University Press, 2011 DOI: 10.1093/acprof:oso/9780198566946.001.0001
  • [RS05] Ran Raz and Amir Shpilka “Deterministic polynomial identity testing in non-commutative models” In Computational Complexity 14.1, 2005, pp. 1–19 DOI: 10.1007/s00037-005-0188-8
  • [VW23] Michèle Vergne and Michael Walter “Moment cone membership for quivers in strongly polynomial time” In arXiv, 2023

Appendix A Elementary proof of King’s criterion

Here, we provide an elementary proof of King’s criterion. In Section A.1, we start from the base case where α\alpha is the all-one vector, i.e., V(a)V(a)\in\mathbb{C} for all arcs aQ1a\in Q_{1}.

A.1 The case when α=𝟏\alpha=\mathbf{1}

The following lemma characterizes the σ\sigma-semistability of a representation with α=𝟏\alpha=\mathbf{1}, which has already been mentioned in Section 1.4. Here, we show this via linear programming. Recall that the support quiver of a representation VV of a quiver QQ means the subquiver of QQ whose arc set is supp(V){aQ1:V(a)0}\operatorname{supp}(V)\coloneqq\{a\in Q_{1}:V(a)\neq 0\}.

Lemma A.1.

Let VV be a representation of QQ with dimension vector α=𝟏\alpha=\mathbf{1} and σ\sigma a weight. Then, the following are equivalent.

  1. (1)

    VV is σ\sigma-semistable.

  2. (2)

    infxQ0hV,σ(x)>0\inf_{x\in\mathbb{R}^{Q_{0}}}h_{V,\sigma}(x)>0, where

    hV,σ(x)aQ1|V(a)|2ex(ha)x(ta)+exp(iQ0σ(i)x(i))(xQ0).\displaystyle h_{V,\sigma}(x)\coloneqq\sum_{a\in Q_{1}}\lvert V(a)\rvert^{2}e^{x(ha)-x(ta)}+\exp\left(\sum_{i\in Q_{0}}\sigma(i)x(i)\right)\quad(x\in\mathbb{R}^{Q_{0}}). (A.1)
  3. (3)

    There exists an integral flow φ+supp(V)\varphi\in\mathbb{Z}_{+}^{\operatorname{supp}(V)} with φ=σ\partial\varphi=\sigma on the support quiver of VV.

  4. (4)

    σ(Q0)=0\sigma(Q_{0})=0 and σ(X)0\sigma(X)\leq 0 for each lower set XX of the support quiver of VV.

Proof.

By definition, VV is σ\sigma-semistable if and only if

infgGL(Q,α)(aQ1|ghaV(a)gta1|2+|χσ(g)|2)\displaystyle\inf_{g\in\operatorname{GL}(Q,\alpha)}\left(\sum_{a\in Q_{1}}\lvert g_{ha}V(a)g_{ta}^{-1}\rvert^{2}+\lvert\chi_{\sigma}(g)\rvert^{2}\right) (A.2)
=infgGL(Q,α)(aQ1|V(a)|2|gha|2|gta|2+iQ0|gi|2σ(i))>0.\displaystyle=\inf_{g\in\operatorname{GL}(Q,\alpha)}\left(\sum_{a\in Q_{1}}\lvert V(a)\rvert^{2}\lvert g_{ha}\rvert^{2}\lvert g_{ta}\rvert^{-2}+\prod_{i\in Q_{0}}\lvert g_{i}\rvert^{2\sigma(i)}\right)>0. (A.3)

Letting |gi|2=ex(i)\lvert g_{i}\rvert^{2}=e^{x(i)}, where x(i)x(i)\in\mathbb{R}, we can rewrite this as infxQ0hV,σ(x)>0\inf_{x\in\mathbb{R}^{Q_{0}}}h_{V,\sigma}(x)>0. This is equivalent to the optimal value of the LP

min\displaystyle\min iQ0σ(i)x(i)\displaystyle\quad\sum_{i\in Q_{0}}\sigma(i)x(i) (A.4)
s.t. x(ha)x(ta)1(aQ1,V(a)0)\displaystyle\quad x(ha)-x(ta)\leq-1\quad(a\in Q_{1},V(a)\neq 0) (A.5)

being not equal to -\infty. Note that the above LP is feasible because QQ is acyclic. By the LP duality theorem, this is equivalent to the existence of a feasible solution of the dual LP

max\displaystyle\max asupp(V)φ(a)\displaystyle\quad-\sum_{a\in\operatorname{supp}(V)}\varphi(a) (A.6)
s.t. φ=σ\displaystyle\quad\partial\varphi=\sigma (A.7)
φ+supp(V).\displaystyle\quad\varphi\in\mathbb{R}_{+}^{\operatorname{supp}(V)}. (A.8)

In particular, we can equivalently replace φ+supp(V)\varphi\in\mathbb{R}_{+}^{\operatorname{supp}(V)} with φ+supp(V)\varphi\in\mathbb{Z}_{+}^{\operatorname{supp}(V)} since the constraint φ=σ\partial\varphi=\sigma is totally dual integral. By Gale’s theorem [Gal57] (see, e.g., [KV18, Theorem 9.2]), this is equivalent to σ(Q0)=0\sigma(Q_{0})=0 and σ(X)0\sigma(X)\leq 0 for each lower set XX of the support quiver of VV. ∎

For later use, we give a lower bound on infxQ0hV,σ(x)\inf_{x\in\mathbb{R}^{Q_{0}}}h_{V,\sigma}(x) that is continuous in VV. Let Φσ+Q1\Phi^{\sigma}\subseteq\mathbb{Z}_{+}^{Q_{1}} be the set of all integral flows on QQ with φ=σ\partial\varphi=\sigma. Note that Φσ\Phi^{\sigma} is a finite set because QQ is acyclic.

Lemma A.2.

If Φσ\Phi^{\sigma}\neq\emptyset, we have

infxQ0hV,σ(x)1|Φσ|φΦσasupp(φ)(|V(a)|2φ(a))φ(a)1+φ1.\displaystyle\inf_{x\in\mathbb{R}^{Q_{0}}}h_{V,\sigma}(x)\geq\frac{1}{|\Phi^{\sigma}|}\sum_{\varphi\in\Phi^{\sigma}}\prod_{a\in\operatorname{supp}(\varphi)}\left(\frac{|V(a)|^{2}}{\varphi(a)}\right)^{\frac{\varphi(a)}{1+\lVert\varphi\rVert_{1}}}. (A.9)
Proof.

We fix φΦσ\varphi\in\Phi^{\sigma} and show

infxQ0hV,σ(x)asupp(φ)(|V(a)|2φ(a))φ(a)1+φ1.\displaystyle\inf_{x\in\mathbb{R}^{Q_{0}}}h_{V,\sigma}(x)\geq\prod_{a\in\operatorname{supp}(\varphi)}\left(\frac{|V(a)|^{2}}{\varphi(a)}\right)^{\frac{\varphi(a)}{1+\lVert\varphi\rVert_{1}}}. (A.10)

If V(a)=0V(a)=0 for some asupp(φ)a\in\operatorname{supp}(\varphi), the bound trivially holds as hV,σ(x)0h_{V,\sigma}(x)\geq 0. Suppose V(a)0V(a)\neq 0 for all asupp(φ)a\in\operatorname{supp}(\varphi). By φ=σ\partial\varphi=\sigma, we have

iQ0σ(i)x(i)=iQ0(φ)(i)x(i)=aQ1φ(a)(x(ta)x(ha))\displaystyle\sum_{i\in Q_{0}}\sigma(i)x(i)=\sum_{i\in Q_{0}}(\partial\varphi)(i)x(i)=\sum_{a\in Q_{1}}\varphi(a)(x(ta)-x(ha)) (A.11)

for xQ0x\in\mathbb{R}^{Q_{0}}. Thus,

infxQ0hV,σ(x)\displaystyle\inf_{x\in\mathbb{R}^{Q_{0}}}h_{V,\sigma}(x) =infxQ0(aQ1|V(a)|2ex(ha)x(ta)+exp(aQ1φ(a)(x(ta)x(ha))))\displaystyle=\inf_{x\in\mathbb{R}^{Q_{0}}}\left(\sum_{a\in Q_{1}}\lvert V(a)\rvert^{2}e^{x(ha)-x(ta)}+\exp\left(\sum_{a\in Q_{1}}\varphi(a)(x(ta)-x(ha))\right)\right) (A.12)
infyQ1(aQ1|V(a)|2ey(a)+exp(aQ1φ(a)y(a)))\displaystyle\geq\inf_{y\in\mathbb{R}^{Q_{1}}}\left(\sum_{a\in Q_{1}}\lvert V(a)\rvert^{2}e^{-y(a)}+\exp\left(\sum_{a\in Q_{1}}\varphi(a)y(a)\right)\right) (A.13)
=infysupp(φ)(asupp(φ)|V(a)|2ey(a)+exp(asupp(φ)φ(a)y(a))).\displaystyle=\inf_{y\in\mathbb{R}^{\operatorname{supp}(\varphi)}}\left(\sum_{a\in\operatorname{supp}(\varphi)}\lvert V(a)\rvert^{2}e^{-y(a)}+\exp\left(\sum_{a\in\operatorname{supp}(\varphi)}\varphi(a)y(a)\right)\right). (A.14)

As (A.14) is the minimization of a strictly convex function, the unique stationary point ysupp(φ)y^{*}\in\mathbb{R}^{\operatorname{supp}(\varphi)} attains the minimum if it exists. Letting

g(y)exp(asupp(φ)φ(a)y(a))=asupp(φ)eφ(a)y(a),\displaystyle g(y)\coloneqq\exp\left(\sum_{a\in\operatorname{supp}(\varphi)}\varphi(a)y(a)\right)=\prod_{a\in\operatorname{supp}(\varphi)}e^{\varphi(a)y(a)}, (A.15)

we can write the first-order optimality criterion for yy^{*} as |V(a)|2ey(a)+φ(a)g(y)=0-|V(a)|^{2}e^{-y^{*}(a)}+\varphi(a)g(y^{*})=0 for asupp(φ)a\in\operatorname{supp}(\varphi). Thus, we have

y(a)=log|V(a)|2φ(a)g(y).\displaystyle y^{*}(a)=\log\frac{|V(a)|^{2}}{\varphi(a)g(y^{*})}. (A.16)

Substituting this back to g(y)g(y), we obtain

g(y)=asupp(φ)(|V(a)|2φ(a)g(y))φ(a)=g(y)φ1asupp(φ)(|V(a)|2φ(a))φ(a)\displaystyle g(y^{*})=\prod_{a\in\operatorname{supp}(\varphi)}\left(\frac{|V(a)|^{2}}{\varphi(a)g(y^{*})}\right)^{\varphi(a)}=g(y^{*})^{-\lVert\varphi\rVert_{1}}\prod_{a\in\operatorname{supp}(\varphi)}\left(\frac{|V(a)|^{2}}{\varphi(a)}\right)^{\varphi(a)} (A.17)
and
g(y)=asupp(φ)(|V(a)|2φ(a))φ(a)1+φ1,\displaystyle g(y^{*})=\prod_{a\in\operatorname{supp}(\varphi)}\left(\frac{|V(a)|^{2}}{\varphi(a)}\right)^{\frac{\varphi(a)}{1+\lVert\varphi\rVert_{1}}}, (A.18)

implying the desired lower bound as (A.14) is at least g(y)g(y^{*}). ∎

A.2 General case

Now, let us move on to the general case to show King’s criterion. Let VV be a representation with the dimension vector α\alpha. By definition, VV is σ\sigma-semistable if and only if

infgGL(Q,α)(aQ1ghaV(a)gta1F2+|χσ(g)|2)\displaystyle\inf_{g\in\operatorname{GL}(Q,\alpha)}\left(\sum_{a\in Q_{1}}\left\lVert g_{ha}V(a)g_{ta}^{-1}\right\rVert_{F}^{2}+\lvert\chi_{\sigma}(g)\rvert^{2}\right) (A.19)
=infgGL(Q,α)(aQ1tr(V(a)ghaghaV(a)(gtagta)1)+iQ0det(gigi)σ(i))\displaystyle=\inf_{g\in\operatorname{GL}(Q,\alpha)}\left(\sum_{a\in Q_{1}}\operatorname{tr}(V(a)^{\dagger}g_{ha}^{\dagger}g_{ha}V(a)(g_{ta}^{\dagger}g_{ta})^{-1})+\prod_{i\in Q_{0}}\det(g_{i}^{\dagger}g_{i})^{\sigma(i)}\right) (A.20)
=infXPD(Q,α)(aQ1tr(V(a)XhaV(a)Xta1)+iQ0detXiσ(i))>0.\displaystyle=\inf_{X\in\operatorname{PD}(Q,\alpha)}\left(\sum_{a\in Q_{1}}\operatorname{tr}(V(a)^{\dagger}X_{ha}V(a)X_{ta}^{-1})+\prod_{i\in Q_{0}}\det X_{i}^{\sigma(i)}\right)>0. (A.21)

Here, PD(Q,α)iQ0PD(α(i))\operatorname{PD}(Q,\alpha)\coloneqq\prod_{i\in Q_{0}}\operatorname{PD}(\alpha(i)), in which PD(n)\operatorname{PD}(n) denotes the set of positive definite matrices of degree nn. Let us take an eigendecomposition of XiX_{i} as

Xi=UiDiag(exi)Ui=j[α(i)]exi(j)𝐮i,j𝐮i,j,\displaystyle X_{i}=U_{i}\operatorname{Diag}(e^{x_{i}})U_{i}^{\dagger}=\sum_{j\in[\alpha(i)]}e^{x_{i}(j)}\mathbf{u}_{i,j}\mathbf{u}_{i,j}^{\dagger}, (A.22)

where xiα(i)x_{i}\in\mathbb{R}^{\alpha(i)}, UiU(α(i))U_{i}\in U(\alpha(i)), and 𝐮i,j\mathbf{u}_{i,j} is the jjth column of UiU_{i}. Then, the objective function inside the infimum becomes

f(x,U)aQ1j[α(ha)]k[α(ta)]|𝐮ha,jV(a)𝐮ta,k|2exha(j)xta(k)+exp(iQ0σ(i)j[α(i)]xi(j)).\displaystyle f(x,U)\coloneqq\sum_{a\in Q_{1}}\sum_{j\in[\alpha(ha)]}\sum_{k\in[\alpha(ta)]}{\bigl{|}\mathbf{u}_{ha,j}^{\dagger}V(a)\mathbf{u}_{ta,k}\bigr{|}}^{2}e^{x_{ha}(j)-x_{ta}(k)}+\exp\left(\sum_{i\in Q_{0}}\sigma(i)\sum_{j\in[\alpha(i)]}x_{i}(j)\right). (A.23)

Observe that f(x,U)f(x,U) for a fixed UU is in the form of the objective function (A.1) of the base case. More precisely, let Q=(Q0,Q1)Q^{\prime}=(Q_{0}^{\prime},Q_{1}^{\prime}) be a quiver such that each vertex iQ0i\in Q_{0} is copied into α(i)\alpha(i) copies (i,j)(i,j) for j[α(i)]j\in[\alpha(i)] and each arc aQ1a\in Q_{1} is copied α(ha)α(ta)\alpha(ha)\alpha(ta) times to connect the copies of the original endpoints. Then, V|U={𝐮ha,jV(a)𝐮ta,k:aQ1,j[α(ha)],k[α(ta)]}V|_{U}=\bigl{\{}\mathbf{u}_{ha,j}^{\dagger}V(a)\mathbf{u}_{ta,k}:a\in Q_{1},\ j\in[\alpha(ha)],\ k\in[\alpha(ta)]\bigr{\}} is a representation of QQ^{\prime} with all-one dimension vector. Let σ\sigma^{\prime} be a weight on QQ^{\prime} such that σ(i,j)=σ(i)\sigma^{\prime}(i,j)=\sigma(i). Then, we have f(x,U)=hV|U,σ(x)f(x,U)=h_{V|_{U},\sigma^{\prime}}(x).

Since infx,Uf(x,U)>0\inf_{x,U}f(x,U)>0 implies infxf(x,U)>0\inf_{x}f(x,U)>0 for any fixed unitary UU, by Lemma A.1, the σ\sigma-semistability of VV implies the following property:

  • (P)

    for any unitary UU, σ(Q0)=0\sigma^{\prime}(Q_{0})=0 and σ(X)0\sigma^{\prime}(X)\leq 0 for each lower set XX of the support quiver of V|UV|_{U}.

To show the converse direction, suppose that (P) holds. By Lemma A.1, this implies the existence of φΦσ\varphi\in\Phi^{\sigma^{\prime}} with supp(φ)supp(V|U)\operatorname{supp}(\varphi)\subseteq\operatorname{supp}(V|_{U}) for any UU. By Lemma A.2, infxf(x,U)\inf_{x}f(x,U) is lower bounded by some positive-valued continuous function of UU. Since the direct sum of the unitary groups is compact, we have that infx,Uf(x,U)>0\inf_{x,U}f(x,U)>0, i.e., VV is σ\sigma-semistable.

We finally show that (P) is equivalent to King’s criterion. First, note that

σ(Q0)=iQ0σ(i)α(i)=σ(α)=σ(dim¯V).\displaystyle\sigma^{\prime}(Q_{0}^{\prime})=\sum_{i\in Q_{0}}\sigma(i)\alpha(i)=\sigma(\alpha)=\sigma(\operatorname{\underline{dim}}V). (A.24)

Thus, σ(dim¯V)=0\sigma(\operatorname{\underline{dim}}V)=0 if and only if σ(Q0)=0\sigma^{\prime}(Q_{0}^{\prime})=0. Let XX be a lower set of the support quiver of V|UV|_{U}. Then, W(i)={𝐮i,j:(i,j)X}W(i)=\langle\{\mathbf{u}_{i,j}:(i,j)\in X\}\rangle (iQ0i\in Q_{0}) defines a subrepresentation of VV and σ(dim¯W)=σ(X)0\sigma(\operatorname{\underline{dim}}W)=\sigma^{\prime}(X)\leq 0. Conversely, for a subrepresentation WW, take a unitary UU such that W(i)W(i) is spanned by 𝐮i,j\mathbf{u}_{i,j} for j[dimW(i)]j\in[\dim W(i)]. Define XQ0X\subseteq Q_{0}^{\prime} as X={(i,j)Q0:𝐮i,jW(i)}X=\{(i,j)\in Q_{0}^{\prime}:\mathbf{u}_{i,j}\in W(i)\}. Then, XX is a lower set of the support quiver of V|UV|_{U} and σ(X)=σ(dim¯W)0\sigma^{\prime}(X)=\sigma(\operatorname{\underline{dim}}W)\leq 0. This completes the proof of King’s criterion.