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

On independent GKM-graphs without nontrivial extensions

Grigory Solomadin Laboratory of algebraic topology and its applications, Faculty of computer science, National Research University Higher School of Economics, Russian Federation [email protected]
Abstract.

In this paper an example of a kk-independent (n,k)(n,k)-type GKM-graph without nontrivial extensions is constructed for any nk3n\geq k\geq 3. It is shown that this example cannot be realized by a GKM-manifold for any n=k=3n=k=3 or nk4n\geq k\geq 4.

Key words and phrases:
Torus action, GKM-theory
The research was partially supported by the “RUDN University program 5–100” program. The article was prepared within the framework of the HSE University Basic Research Program

1. Introduction

GKM-theory [gkm-98], [gu-za-01] provides with a useful method for computation of the equivariant cohomology ring for a wide class of manifolds equipped with a torus action (having only isolated fixed points) called GKM-manifolds. The main tool of this theory is a graph that is naturally asssociated with the orbit space of the equivariant 11-skeleton for a GKM-manifold. The labels on the edges of this graph are given by the tangent weights of the fixed points associated with the torus action. By abstracting from the torus action one obtains the definition of a GKM-graph axiomatized in [gu-za-01]. Apart from the equivariant cohomology ring many other important geometric objects and topological invariants related to a GKM-manifold can be studied in terms of the respective GKM-graph, such as Betti numbers [gu-za-01], invariant almost complex structures [gu-ho-za-06], invariant symplectic and Kähler structures [go-ko-zo-20], etc.

A GKM-manifold M2nM^{2n} is called a GKM-manifold in jj-general position if any jj tangent weights at any fixed point of this action on M2nM^{2n} are linearly independent. A jj-independent GKM-graph is defined similarly. Properties of a GKM-action of a TkT^{k}-action on M2nM^{2n} in jj-general position were studied in several papers (in a wider context of equivariantly formal smooth actions with isolated fixed points). The number nkn-k is called the complexity of the TkT^{k}-action on M2nM^{2n}. In [ma-pa-06] it was shown that the equivariant cohomology ring of a torus manifold with vanishing odd cohomology groups (that is, a 2n2n-dimensional GKM-manifold of complexity 0 which is automatically nn-independent) is isomorphic to the Stanley-Reisner ring of the face poset of the GKM-graph. Also see [ma-ma-pa-07] for treatment of equivariant topology for torus manifolds from GKM-perspective. Furthermore, any GKM-manifold M2nM^{2n} of complexity 11 with a GKM-action in general position (that is, in (n1)(n-1)-general position) has a description [ay-ma-so-22] of the equivariant cohomology ring in terms of face rings similar to [ma-pa-06]. The poset of faces SMS_{M} in the orbit space M/TM/T was associated to a GKM-manifold MM with the TT-action in [ay-ma-so-22]. We remark that in the case of complexity 0 this object was studied in earlier papers [ma-pa-06] (for a locally standard action, so that M/TM/T is a manifold with corners) and in [ay-ma-19] (for a general case, so that M/TM/T is a homological cell complex). It was shown in [ay-ma-so-22] that for a GKM-manifold in jj-general position the subposets (SM)<s(S_{M})_{<s} and (SM)r(S_{M})_{r} are min{dims1,j+1}\min\{\dim s-1,j+1\}- and min{r1,j+1}\min\{r-1,j+1\}-acyclic, respectively, for any r>0r>0 and sSMs\in S_{M}. This implies by [ma-pa-06, ay-ma-19] that the corresponding orbit space M/TM/T of MM is (j+1)(j+1)-acyclic. This result was proved in [ma-pa-06] for j=nj=n (nn-independent case) and in [ay-ma-19] for the case of arbitrary jj. On the other hand, by dropping the condition of general position, an arbitrary Tn1T^{n-1}-action on a smooth manifold M2nM^{2n} (of complexity 11) with isolated fixed points shows much more complicated behaviour of the orbit space homology [ay-ch-19]. Therefore, GKM-actions in jj-general position form a particularly nice class of GKM-actions exhibiting many useful properties (depending on the value of jj).

A natural question (called an extension problem in [ku-19]) is how to determine whether a given GKM-action extends to an effective GKM-action of a torus of greater dimension (so that the complexity of the new action is lower) on the same manifold, or not. This question leads to a search of a kk-independent (n,k)(n,k)-type GKM-action that is nonextendible for arbitrary nk2n\geq k\geq 2. Homogeneous GKM-manifolds [gu-ho-za-06] supply with many interesting examples of GKM-graphs, including those of positive complexity. Recall that if the Euler characteristic of a homogeneous complex manifold G/HG/H is nonzero then it is a GKM-manifold by [gu-ho-za-06], where GG is a compact connected semisimple Lie group, HH is an arbitrary parabolic subgroup in GG and TT is any maximal torus of HGH\subset G. Two phenomena occur for homogeneous GKM-manifolds related to the search problem mentioned above. Firstly, S. Kuroki proved the following gap property for GG of type AnA_{n} with the natural GKM TT-action on G/HG/H as above (unpublished): if the respective GKM-graph is jj-independent for some j4j\geq 4 then it is nn-independent, where nn is the rank of GG. This leads to a conjecture that any homogeneous GKM-manifold satisfies such a gap property. Secondly, the natural Tn1T^{n-1}-action on the Grassmanian Grk(n)Gr_{k}(\mathbb{C}^{n}) of kk-planes in n\mathbb{C}^{n} is a GKM-action admitting no nonisomorphic extensions for any kk. This can be shown by modifying the proof for k=2k=2 from [ku-19] in a simple way. This result leads to another conjecture that most of the homogeneous GKM-manifolds are nonextendible with possibly a few exceptions determined by the condition that the respective automorphism group identity component Aut0(G/H)\operatorname{Aut}^{0}(G/H) is greater than GG. (The description of Aut0(G/H)\operatorname{Aut}^{0}(G/H) is given in [ak-95], for instance.)

Therefore the extension problem seems to be a bit challenging. Instead, one can study a rather simplistic extension problem for arbitrary GKM-graphs by replacing a GKM-action of TkT^{k} on M2nM^{2n} with an (n,k)(n,k)-type kk-independent GKM-graph Γ\Gamma (not necessarily coming from a GKM-action). An occurring new realization problem here is to determine whether a given GKM-graph Γ\Gamma is a GKM-graph of some GKM-manifold. We remark that some results on realization problem were obtained in dimensions n=1,2n=1,2 in [ca-ga-ka] for GKM-graphs (in terms of conditions given by the ABBV localization formula), and for GKM-orbifolds in dimension n=2n=2 in [da-ku-so-18, Remark 4.13].

In this paper, we give new examples of non-realizable and non-extendible actions of different complexity. The precise statement is given in the following main theorem of this paper (see Theorem LABEL:thm:example_).

Theorem 1.

For any nk3n\geq k\geq 3 there exists an (n,k)(n,k)-type GKM-graph Γ\Gamma satisfying the following properties:

  1. (i)

    The GKM-graph Γ\Gamma is kk-independent;

  2. (ii)

    The GKM-graph Γ\Gamma has no nontrivial extensions;

  3. (iii)

    The GKM-graph Γ\Gamma is not realized by a GKM-action if n=k=3n=k=3 or nk4n\geq k\geq 4.

The necessary tools for the proof of Theorem 1 are twofold. Firstly, in Section 2 we give a sufficient criterion (Corollary 2.12) of non-extendibility of any GKM-graph satisfying some condition formulated in terms of chords for a face in the GKM-graph. Secondly, in Section 3 for any (j+1)(j+1)-independent GKM-manifold MM with the corresponding GKM-graph Γ\Gamma, we compare (Proposition 3.9) some subposets in the face poset SΓS_{\Gamma} of a GKM-graph Γ\Gamma with the subposets in the face poset SMS_{M} of MM. The obtained comparison implies (by [ay-ma-so-22]) partial acyclicity for some subposets in SΓS_{\Gamma}. In this case (SΓ)Ξop(S_{\Gamma})_{\leq\Xi}^{op} turns out to be a simplicial poset for any face Ξ\Xi of dimension not exceeding jj in Γ\Gamma (Proposition 3.14). The respective Euler characteristic of the order complex for the poset (SΓ)<Ξop(S_{\Gamma})_{<\Xi}^{op} is computed in terms of the corresponding ff-numbers (Proposition 3.13).

In Section 4 we explicitly construct an example which is suitable for the proof of Theorem 1. Here is a brief outline of the construction. Firstly we introduce an infinite (d+1)(d+1)-regular graph embedded into d\mathbb{R}^{d}. We endow it with an axial function in order to obtain a torus graph Γ(d,0)\Gamma(d,0). We add to it edges and define the axial function agreeing to that of Γ(d,0)\Gamma(d,0) causing the increase of the complexity for a GKM-graph. The resulting (d+1+r,d+1)(d+1+r,d+1)-type GKM graph Γ(d,r)\Gamma(d,r) is (d+1)(d+1)-independent. We remark that the definition of the axial function on Γ(d,r)\Gamma(d,r) is obtained by applying the decision method of Tarski [ta-51] to some Vandermonde matrices (Lemma 4.9) and therefore it is implicit. The infinite GKM graph Γ(d,r)\Gamma(d,r) is periodic (invariant) with respect to the subgroup 2r+1dd2^{r+1}\cdot\mathbb{Z}^{d}\subset\mathbb{Z}^{d} in the group of parallel translations in d\mathbb{R}^{d}. The quotient Γd+1d+1+r:=Γ2r+1(d,r)\Gamma_{d+1}^{d+1+r}:=\Gamma^{2^{r+1}}(d,r) of the GKM-graph Γ(d,r)\Gamma(d,r) by the group 2r+1d2^{r+1}\cdot\mathbb{Z}^{d} is shown to be a well-defined GKM-graph without multiple edges and loops.

In the concluding Section LABEL:sec:proof of this paper we prove that Γd+1d+1+r\Gamma_{d+1}^{d+1+r} satisfies all conditions of Theorem 1 (where we put n=d+1+rn=d+1+r, k=n+1k=n+1). We apply the results on acyclicity of face subposets in GKM-manifolds (in the case of complexity 0) from [ma-pa-06] in order to prove non-realizability of the constructed GKM-graph Γd+1d+1+r\Gamma_{d+1}^{d+1+r} by studying the respective Euler characteristic (by the comparison results mentioned above). This argument relies on the explicit computation of face numbers in the torus graph Γa(d,0)\Gamma^{a}(d,0) (Lemma LABEL:lm:eucomp). The nonextendibility is proved by using the method of chords mentioned above.

2. An obstruction to a GKM-graph extension

In this section we recall some definitions from GKM-theory (we follow the notation from [gu-za-01] and [ku-09]). We introduce an obstruction to have extensions for a given GKM-graph in terms of chords.

Definition 2.1.

[gu-za-01] A GKM-graph Γ\Gamma is a triple ((V,E),,α)((V,E),\nabla,\alpha) consisting of:

  • a graph (V,E)(V,E) with the set of vertices VV\neq\varnothing and with the set of edges EV×VΔ(V)E\subseteq V\times V\setminus\Delta(V), where Δ(V):={(v,v)|vV}V×V\Delta(V):=\{(v,v)|\ v\in V\}\subseteq V\times V and it is required that for an edge e=(u,v)Ee=(u,v)\in E one has e¯:=(v,u)E\overline{e}:=(v,u)\in E;

  • a collection of bijections

    ={e:starΓi(e)starΓt(e)},\nabla=\{\nabla_{e}\colon\operatorname{star}_{\Gamma}i(e)\to\operatorname{star}_{\Gamma}t(e)\},

    such that the identity

    e(e)=e¯,\nabla_{e}(e)=\overline{e},

    holds for any eEe\in E, where starΓi(e):={eE|i(e)=v}\operatorname{star}_{\Gamma}i(e):=\{e\in E|\ i(e)=v\} is the star of Γ\Gamma at i(e)i(e);

  • a function α:Ek\alpha\colon E\to\mathbb{Z}^{k} satisfying the rank condition

    αΓ:=α(e)|estarΓ(v)=k,\alpha\langle{\Gamma}\rangle:=\mathbb{Z}\langle\alpha(e)|\ e\in\operatorname{star}_{\Gamma}(v)\rangle=\mathbb{Z}^{k},

    the opposite sign condition

    α(e¯)=α(e),\alpha(\overline{e})=-\alpha(e),

    and the congruence condition

    α(e(e))=α(e)+ce(e)α(e),\alpha(\nabla_{e}(e^{\prime}))=\alpha(e^{\prime})+c_{e}(e^{\prime})\alpha(e),

    for some integer ce(e)c_{e}(e^{\prime})\in\mathbb{Z} and for any e,eEe,e^{\prime}\in E with a common source, and vVv\in V.

The collection \nabla is called a connection of Γ\Gamma and the function α\alpha is called an axial function of Γ\Gamma.

In this section we fix a GKM-graph Γ\Gamma with the corresponding connected nn-valent graph (V,E)=(VΓ,EΓ)(V,E)=(V_{\Gamma},E_{\Gamma}), axial function α:Ek\alpha\colon E\to\mathbb{Z}^{k} on Γ\Gamma and a connection \nabla on it.

Definition 2.2.

[gu-za-01] A connected rr-regular subgraph Ξ\Xi of Γ\Gamma is called an rr-face (or a face) of Γ\Gamma, if e(e)EΞ\nabla_{e}(e^{\prime})\in E_{\Xi} holds for any e,eEΞe,e^{\prime}\in E_{\Xi} such that i(e)=i(e)i(e)=i(e^{\prime}) (in [gu-za-01] it is called a totally geodesic subgraph). Any edge estarΓvstarΞve\in\operatorname{star}_{\Gamma}v\setminus\operatorname{star}_{\Xi}v is called a transversal edge to a face Ξ\Xi in Γ\Gamma, where vVΞv\in V_{\Xi}. Let

αΞ=αvΞ:=α(e)|estarΞ(v)k,\alpha\langle\Xi\rangle=\alpha_{v}\langle\Xi\rangle:=\mathbb{Z}\langle\alpha(e)|\ e\in\operatorname{star}_{\Xi}(v)\rangle\subseteq\mathbb{Z}^{k},

be a span of a face Ξ\Xi in Γ\Gamma, where α(e)|estarΞ(v)\mathbb{Z}\langle\alpha(e)|\ e\in\operatorname{star}_{\Xi}(v)\rangle denotes the \mathbb{Z}-linear span of vectors α(e)\alpha(e), where ee runs over starΞ(v)\operatorname{star}_{\Xi}(v). The GKM-graph Γ\Gamma is called a GKM-graph of (n,k)(n,k)-type, if Γ\Gamma is nn-valent and the rank rkα:=rkαΓ\operatorname{rk}\alpha:=\operatorname{rk}\alpha\langle\Gamma\rangle of α\alpha is equal to kk.

Remark 2.3.

A face Ξ\Xi of a GKM-graph Γ\Gamma becomes a well-defined GKM-graph by taking restrictions of the connection and of the axial function from Γ\Gamma to Ξ\Xi. The span of the face Ξ\Xi is well defined because of the identity αpΞ=αqΞ\alpha_{p}\langle\Xi\rangle=\alpha_{q}\langle\Xi\rangle for every p,qVΞp,q\in V_{\Xi} which immediately follows from the congruence condition.

Definition 2.4.

A GKM-graph Γ\Gamma is called jj-complete if for any vVv\in V, any integer iji\leq j and any distinct edges e1,,eistarΓ(v)e_{1},\dots,e_{i}\in\operatorname{star}_{\Gamma}(v) there exists an ii-face Ξ\Xi of Γ\Gamma such that starΞ(v)={e1,,ei}\operatorname{star}_{\Xi}(v)=\{e_{1},\dots,e_{i}\} holds. An nn-regular nn-complete GKM-graph is called a complete GKM-graph. A GKM-graph Γ\Gamma is called jj-independent if for any vVv\in V and any distinct edges e1,,ejstarΓve_{1},\dots,e_{j}\in\operatorname{star}_{\Gamma}v the values α(e1),,α(ej)\alpha(e_{1}),\dots,\alpha(e_{j}) of the axial function α\alpha on Γ\Gamma are linearly independent in k\mathbb{Z}^{k}.

Definition 2.5.

Let Ξ\Xi be a face of Γ\Gamma. We call a transversal edge eEΓe\in E_{\Gamma} to Ξ\Xi a chord of the face Ξ\Xi, if i(e),t(e)VΞi(e),t(e)\in V_{\Xi}. If the face Ξ\Xi in Γ\Gamma admits no chords then we call Ξ\Xi a chordless face of Γ\Gamma.

Example 2.6.

The standard T2T^{2}-action on the flag manifold l(3)\mathcal{F}l(3) is a GKM-action with the GKM-graph Γ\Gamma of (3,2)(3,2)-type and with K3,3K_{3,3} as the underlying graph [gu-ho-za-06, p.40]. The GKM-graph Γ\Gamma satisfies the opposite sign condition. It has five 22-faces (three 44-cycles and two 66-cycles). For any such 66-cycle 22-face Ξ\Xi the remaining 33 transversal edges in Γ\Gamma are chords of Ξ\Xi.

Proposition 2.7.

Let Ξ\Xi be a jj-face of a (j+1)(j+1)-complete GKM-graph Γ\Gamma. Then for any chord eEΓe\in E_{\Gamma} of Ξ\Xi and any edge path γ=(e1,,er)\gamma=(e_{1},\dots,e_{r}) in Ξ\Xi such that i(γ)=i(e)i(\gamma)=i(e), t(γ)=t(e)t(\gamma)=t(e), one has Πγe=e¯\Pi_{\gamma}e=\overline{e}, where by definition (see [ta-04])

Πγ:starΓ(i(γ))starΓ(t(γ)),Πγ(e):=ere1(e).\Pi_{\gamma}:\ \operatorname{star}_{\Gamma}(i(\gamma))\to\operatorname{star}_{\Gamma}(t(\gamma)),\quad\Pi_{\gamma}(e):=\nabla_{e_{r}}\circ\dots\circ\nabla_{e_{1}}(e).
Proof.

By the (j+1)(j+1)-completeness condition, there exists a (j+1)(j+1)-face Φ\Phi of Γ\Gamma such that Ξ\Xi and ee belong to Φ\Phi. Notice that Πe(e)=e¯Φ\Pi_{e}(e)=\overline{e}\in\Phi holds by the definition. The definition of invariance also implies that Πe(e)starΦ(t(γ))starΞ(t(γ))\Pi_{e}(e)\in\operatorname{star}_{\Phi}(t(\gamma))\setminus\operatorname{star}_{\Xi}(t(\gamma)) holds. However, starΦ(t(γ))starΞ(t(γ))\operatorname{star}_{\Phi}(t(\gamma))\setminus\operatorname{star}_{\Xi}(t(\gamma)) has cardinality one, since Ξ\Xi and Φ\Phi are jj- and (j+1)(j+1)-faces, respectively. Hence,

starΦ(t(γ))starΞ(t(γ))={e¯},\operatorname{star}_{\Phi}(t(\gamma))\setminus\operatorname{star}_{\Xi}(t(\gamma))=\{\overline{e}\},

holds. Observe that Πγ(e)starΦ(t(γ))starΞ(t(γ))\Pi_{\gamma}(e)\in\operatorname{star}_{\Phi}(t(\gamma))\setminus\operatorname{star}_{\Xi}(t(\gamma)) holds by the definition of invariance, because γ\gamma belongs to Ξ\Xi by the condition. Therefore, we conclude that Πγ(e)=e¯\Pi_{\gamma}(e)=\overline{e} holds. This proves the claim of the proposition. ∎

Proposition 2.8.

For a chord eEΓe\in E_{\Gamma} of a face Ξ\Xi in Γ\Gamma suppose that there exists an edge path γ\gamma in Ξ\Xi such that i(γ)=i(e)i(\gamma)=i(e), t(γ)=t(e)t(\gamma)=t(e) and Πγe=e¯\Pi_{\gamma}e=\overline{e} hold. Then one has 2α(e)αΞ2\alpha(e)\in\alpha\langle\Xi\rangle.

Proof.

Let γ=(e1,,er)\gamma=(e_{1},\dots,e_{r}). Notice that one has Πγi1eiEΞ\Pi_{\gamma_{i-1}}e_{i}\in E_{\Xi}, because Ξ\Xi is a face in Γ\Gamma, where γi:=(e1,,ei)\gamma_{i}:=(e_{1},\dots,e_{i}) and γ0:=i(γ)\gamma_{0}:=i(\gamma), i=1,,ri=1,\dots,r. Hence, α(Πγi1ei)αΞ\alpha(\Pi_{\gamma_{i-1}}e_{i})\in\alpha\langle\Xi\rangle holds for any i=1,,ri=1,\dots,r. One deduces the identity

(2.1) α(Πγe)=α(e)+i=1rcei(Πγi1e)α(Πγi1ei),\alpha(\Pi_{\gamma}e)=\alpha(e)+\sum_{i=1}^{r}c_{e_{i}}(\Pi_{\gamma_{i-1}}e)\cdot\alpha(\Pi_{\gamma_{i-1}}e_{i}),

from the congruence condition. We conclude that

(2.2) α(Πγe)α(e)αΞ,\alpha(\Pi_{\gamma}e)-\alpha(e)\in\alpha\langle\Xi\rangle,

holds. Notice that the identities α(e¯)=α(e)\alpha(\overline{e})=-\alpha(e), α(e¯)=α(Πγe)\alpha(\overline{e})=\alpha(\Pi_{\gamma}e) hold. Together with the inclusion (2.2) this implies the claim of the proposition. ∎

Proposition 2.9.

Suppose that the underlying graph of Γ\Gamma has finitely many vertices. Then, if Γ\Gamma is (j+1)(j+1)-independent, then Γ\Gamma is jj-complete, where jj\in\mathbb{Z}.

Proof.

Fix a nonzero integer sjs\leq j. Let E:={e1,,es}E^{\prime}:=\{e_{1},\dots,e_{s}\} be an ss-element set of some mutually different edges in Γ\Gamma with a common origin vv. In order to prove the claim it is enough to construct an ss-face Ξ\Xi in Γ\Gamma such that the inclusion

(2.3) EEΞ,E^{\prime}\subseteq E_{\Xi},

holds. We give the inductive definition as follows:

Pi+1:=Pi{Πee|e,ePi},P0:=E,i0.P_{i+1}:=P_{i}\cup\{\Pi_{e}e^{\prime}|\ e,e^{\prime}\in P_{i}\},\ P_{0}:=E^{\prime},\ i\geq 0.

By the definition, the filtration P0P1P_{0}\subseteq P_{1}\subseteq\cdots is bounded by the finite set EΓE_{\Gamma} from above. Hence, there exists NN\in\mathbb{N} such that Pi=PNP_{i}=P_{N} holds for any iNi\geq N. Define the subgraph Ξ\Xi in Γ\Gamma by the formulas

VΞ:={i(e)|ePN},EΞ:=PN.V_{\Xi}:=\{i(e)|\ e\in P_{N}\},\ E_{\Xi}:=P_{N}.

The set PNP_{N} is closed under reversion of an edge operation, because Πe(e)=e¯Pi+1\Pi_{e}(e)=\overline{e}\in P_{i+1} holds for any ePie\in P_{i}. By the condition, for any e,ePNe,e^{\prime}\in P_{N} there exists an edge path γΞ\gamma\subseteq\Xi such that i(γ)=i(e)i(\gamma)=i(e) and t(γ)=i(e)t(\gamma)=i(e^{\prime}) holds. Hence, Ξ\Xi is a connected subgraph in Γ\Gamma. It follows from the definition that Ξ\Xi is a face of Γ\Gamma. It remains to show that Ξ\Xi is an ss-face. Assume the contrary. Then there exists estarΞ(v)Ee\in\operatorname{star}_{\Xi}(v)\setminus E^{\prime}. It follows from the definition that there exist i=1,,si=1,\dots,s and γΞ\gamma\subseteq\Xi such that i(γ)=t(γ)=vi(\gamma)=t(\gamma)=v and

Πγ(e)=e,\Pi_{\gamma}(e^{\prime})=e,

holds. It follows from the formula (2.1) that

α(Πγ(e))α(ej)|j=1,,s.\alpha(\Pi_{\gamma}(e^{\prime}))\in\mathbb{Z}\langle\alpha(e_{j})|\ j=1,\dots,s\rangle.

Hence, the collection of s+1s+1 vectors α(e),α(ej)\alpha(e),\alpha(e_{j}), j=1,,sj=1,\dots,s, is linearly dependent. However, this contradicts the condition of (j+1)(j+1)-independency of Γ\Gamma, because sqs\leq q. We conclude that Ξ\Xi is an ss-face, which proves the claim of the proposition. ∎

Corollary 2.10.

Suppose that the underlying graph of Γ\Gamma has finitely many vertices. Then, if Γ\Gamma is a (j+2)(j+2)-independent (nn-independent, respectively) GKM-graph for some jj\in\mathbb{Z}, then any rr-face (face, respectively) of Γ\Gamma is chordless, where r=1,,jr=1,\dots,j.

Proof.

Assume the contrary. Then there exist an rr-face Ξ\Xi of Γ\Gamma and its chord ee, where rjr\leq j. One has 2α(e)αΞ2\alpha(e)\notin\alpha\langle\Xi\rangle, because Γ\Gamma is (j+2)(j+2)-independent and Ξ\Xi is rr-regular, where rjr\leq j. By Proposition 2.9, Γ\Gamma is (j+1)(j+1)-complete. Then one can apply Propositions 2.7 and 2.8 in order to obtain 2α(e)αΞ2\alpha(e)\in\alpha\langle\Xi\rangle. This contradiction proves the first claim of the corollary. The proof of the second claim is similar to the proof of the first claim. ∎

Definition 2.11.

Let Γ\Gamma^{\prime}, Γ\Gamma be two GKM-graphs with the same underlying graph (V,E)(V,E), the same connection \nabla, with the axial functions α\alpha^{\prime}, α\alpha taking values in k\mathbb{Z}^{k^{\prime}} and k\mathbb{Z}^{k}, respectively. The GKM-graph Γ\Gamma^{\prime} is called an extension of Γ\Gamma (see [ku-19]), if there exists an epimorhism p:kkp\colon\mathbb{Z}^{k^{\prime}}\to\mathbb{Z}^{k} such that p(α(e))=α(e)p(\alpha^{\prime}(e))=\alpha(e) holds for any eEe\in E. We say that an (n,k)(n,k)-type GKM-graph Γ\Gamma has no nontrivial extensions if for any s>0s>0 it does not admit an extension to an (n,k+s)(n,k+s)-type GKM-graph. (This terminology was proposed by S. Kuroki.)

The next corollary is a principal tool for the proof of Theorem 1.

Corollary 2.12.

Suppose that the underlying graph of Γ\Gamma has finitely many vertices. Then, if Γ\Gamma is a (k+1)(k+1)-complete GKM-graph and there exists a kk-face Ξ\Xi of Γ\Gamma such that any transversal edge eEΓe\in E_{\Gamma} to Ξ\Xi is a chord for Ξ\Xi, then Γ\Gamma has no nontrivial extensions.

Proof.

Suppose that there exists an extension of α\alpha to an axial function α~\widetilde{\alpha} of rank k+sk+s for some s>0s>0. Choose a vertex vVΞv\in V_{\Xi}. Then it follows from the condition by Propositions 2.7 and 2.8 that 2α~(e)α~Ξ2\widetilde{\alpha}(e)\in\widetilde{\alpha}\langle\Xi\rangle holds for any edge estarΓ(v)e\in\operatorname{star}_{\Gamma}(v). Hence, k+s=rkα~=rkα~Ξk+s=\operatorname{rk}\widetilde{\alpha}=\operatorname{rk}\widetilde{\alpha}\langle\Xi\rangle. However, by definition rkα~Ξk\operatorname{rk}\widetilde{\alpha}\langle\Xi\rangle\leq k. This contradiction proves the claim. ∎

3. Face posets of a GKM-graph and of a GKM-manifold

In this section we continue to recall some basic notions of GKM-theory and of the related [ma-ma-pa-07, ay-ma-so-22] posets SMS_{M}, SΓS_{\Gamma} of faces arising from the orbit space M/TM/T and from the GKM-graph Γ\Gamma of a given GKM-manifold MM with the TT-action, respectively. We compare some specific simplicial subposets in SMS_{M}, SΓS_{\Gamma} under assumption of jj-general position for MM. After that we recall the P.Hall formula for the Euler characteristic of an order complex for a finite simplicial poset which is used later in the text.

Definition 3.1.

[ma-ma-pa-07, ay-ma-so-22] For a GKM-graph Γ\Gamma the collection SΓS_{\Gamma} of all faces in Γ\Gamma is called a face poset of the GKM-graph Γ\Gamma with the partial order given by inclusion of faces in Γ\Gamma.

Due to [ma-pa-06, Lemma 2.1] one can give the following definition of a GKM-manifold that is equivalent to the standard one (e.g. see [gu-za-01]).

Definition 3.2.

[gkm-98, gu-za-01, ma-pa-06] A smooth manifold M2nM^{2n} with an effective action of Tk=(S1)kT^{k}=(S^{1})^{k} is called a GKM-manifold if the following conditions hold:

  • the set of TkT^{k}-fixed points MTM^{T} in MM is finite and nonempty;

  • the tangent weights of the TkT^{k}-action at any xMTx\in M^{T} are pairwise linearly independent;

  • all odd cohomology groups of MM vanish, i.e. one has Hodd(M;)=0H^{odd}(M;\ \mathbb{Z})=0.

Remark 3.3.

To any complex GKM-manifold one associates a GKM-graph, e.g. see [ku-09]. We notice that for an arbitrary GKM-manifold the opposite sign condition is in general satisfied only up to a sign. We also remark that it is possible to have loops and multiple edges for a GKM-action. In this paper we restrain from considering such torus actions and we use a restricted definition of a GKM-graph (where it is a simple graph). Let TT^{\prime} and TT be two GKM-actions of tori on the same manifold MM. The action of TT^{\prime} is called an extension of the action TT on MM if there is a group monomorphism π:TT\pi\colon T\to T^{\prime} that is equivariant with respect to these torus actions. In other words, the TT-action is the restriction of the TT^{\prime}-action. The epimorphism

p:kHom(T,S1)Hom(T,S1)k,p\colon\mathbb{Z}^{k^{\prime}}\cong\operatorname{Hom}(T^{\prime},S^{1})\to\operatorname{Hom}(T,S^{1})\cong\mathbb{Z}^{k},

corresponding to π\pi induces the extension of the GKM-graphs Γ\Gamma^{\prime}, Γ\Gamma corresponding to the TT^{\prime}- and the TT-action, respectively.

Example 3.4.

The natural T2T^{2}-action on l3\mathcal{F}l_{3} has no notrivial extensions by proving that for the corresponding GKM-graph by Corollary 2.12 (see Example 2.6). This fact may also be easily obtained by the results of [ku-19], or by studying the automorphism group of the homogeneous space l3\mathcal{F}l_{3} (in a different category of complex-analytic torus actions).

Consider a GKM-action of T=TkT=T^{k} on M=M2nM=M^{2n}.

Definition 3.5.

[ay-ma-so-22] For a smooth TT-action on MM, consider the canonical projection p:MQ:=M/Tp\colon M\to Q:=M/T to the respective orbit space, and let

(3.1) Q0Q1Qk=QQ_{0}\subset Q_{1}\subset\cdots\subset Q_{k}=Q
Qi:=p(Mi),Mi={xM:dimTxi},Q_{i}:=p(M_{i}),\ M_{i}=\{x\in M\colon\dim Tx\leqslant i\},

be the filtration on the orbit space QQ, where TxTx denotes the TT-orbit of xx in MM. The closure of a connected component of QiQi1Q_{i}\setminus Q_{i-1} is called an ii-face (or a face) FF of QQ if it contains at least one fixed point.

Definition 3.6.

[ay-ma-so-22] The poset of faces for the GKM-manifold MM is the poset SMS_{M} of faces of nonnegative dimension ordered by inclusion in the orbit space QQ of the TT-manifold MM.

Recall that a topological space XX is called jj-acyclic if H~i(X)=0\tilde{H}^{i}(X)=0 holds for any iji\leq j, and acyclic, if H~(X)=0\tilde{H}^{*}(X)=0 holds. Let SMS_{M} be the poset of faces for a GKM-manifold MM. Let SΓS_{\Gamma} be the poset of faces of nonnegative dimension (ordered by inclusion) of a GKM-graph Γ\Gamma. We need the following particular case of a theorem from [ay-ma-so-22].

Theorem 3.7.

[ay-ma-so-22, Theorem 1] For any GKM-manifold MM of complexity 0 in nn-general position (that is, k=nk=n holds) the (n1)(n-1)-dimensional poset SMop¯\overline{S_{M}^{op}} is (n2)(n-2)-acyclic, where n2n\geq 2.

Lemma 3.8.

[ay-ma-so-22, p.5, Lemma 2.9] The full preimage MF=p1(F)M_{F}=p^{-1}(F) of any face FQF\subseteq Q is a smooth submanifold in MM called a face submanifold in MM.

The claim of the following proposition is reminiscent to [ay-ma-so-22, Lemma 3.8] (although not quite the same).

Proposition 3.9.

For a GKM-manifold MM in (j+1)(j+1)-general position for some j1j\geq 1 the following claims hold.

(i)(i) For any qjq\leq j, any qq-face Ξ\Xi in Γ\Gamma is an equivariant 11-skeleton of a face submanifold in MM and the GKM-graph Ξ\Xi is a torus graph.

(ii)(ii) The span αΞ\alpha\langle\Xi\rangle of Ξ\Xi splits off as a direct factor in k\mathbb{Z}^{k}.

(iii)(iii) The posets (SM)s(Ξ)(S_{M})_{\leq s(\Xi)} and (SΓ)Ξ(S_{\Gamma})_{\leq\Xi} are isomorphic for any face Ξ\Xi of Γ\Gamma such that dimΞj\dim\Xi\leq j, where s(Ξ)s(\Xi) is the face in MM corresponding to Ξ\Xi by (i)(i).

Proof.

Choose vVΞv\in V_{\Xi} and let starΞ(v)={e1,,eq}\operatorname{star}_{\Xi}(v)=\{e_{1},\dots,e_{q}\}. Let GG be a closed subgroup in TT corresponding to the sublattice L:=αi|i=1,,qL:=\mathbb{Z}\langle\alpha_{i}|\ i=1,\dots,q\rangle in k\mathbb{Z}^{k}, where αi:=α(ei)\alpha_{i}:=\alpha(e_{i}). Let G0G_{0} be the identity component (in particular, a torus) of GG. Notice that the sublattice L0kL_{0}\subset\mathbb{Z}^{k} corresponding to the subtorus G0TG_{0}\subset T splits off as a direct factor in k\mathbb{Z}^{k} and that there is a lattice embedding LL0L\subseteq L_{0} of a finite index. The connected component YY of MG0M^{G_{0}} such that vYv\in Y is a smooth manifold with effective TT^{\prime}-action by Lemma 3.8, where T:=T/G0T^{\prime}:=T/G_{0}. Notice that YTMTY^{T^{\prime}}\subseteq M^{T}. The set of weights β1,,βr\beta_{1},\dots,\beta_{r} of the TT^{\prime}-action on YY at vv embed to the set of weights of the TT-action on MM at vv. One has βj|j=1,,r=L0\mathbb{Z}\langle\beta_{j}|\ j=1,\dots,r\rangle=L_{0}. Hence, βj,α1,,αq\beta_{j},\alpha_{1},\dots,\alpha_{q} are linearly dependent for any j=1,,rj=1,\dots,r. Then by linear independence condition we conclude that {β1,,βr}={α1,,αq}\{\beta_{1},\dots,\beta_{r}\}=\{\alpha_{1},\dots,\alpha_{q}\} holds. Therefore, dimY=2q\dim Y=2q, YY is a GKM-manifold, its equivariant 11-skeleton Y1Y_{1} is a GKM-graph Ξ\Xi of type (q,q)(q,q) and Ξ\Xi is a face of the GKM-graph Γ\Gamma. Notice that this implies L=L0L=L_{0}. Hence, the claims (i)(i), (ii)(ii) are proved. By the definition, one has (SM)s(SΓ)s(S_{M})_{\leq s}\subseteq(S_{\Gamma})_{\leq s}. The inverse inclusion holds by (i)(i). This proves (iii)(iii). The proof is complete. ∎

Let PP be a finite poset [st-86]. Recall the following definitions.

Definition 3.10.

[st-86] The order complex of a finite poset PP is the simplicial complex

Δ(P):={σ={I1,I2,,Ik+1}2P|I1<I2<<Ik+1,k0},\Delta(P):=\{\sigma=\{I_{1},I_{2},\dots,I_{k+1}\}\in 2^{P}|\ I_{1}<I_{2}<\cdots<I_{k+1},\ k\geq 0\},

on the vertex set PP consisting of chains of increasing elements in PP. By definition the qq-faces of a simplex σ\sigma are Ii1<Ii2<<Iiq+1I_{i_{1}}<I_{i_{2}}<\cdots<I_{i_{q+1}}, where 1i1<<iq+1k1\leq i_{1}<\cdots<i_{q+1}\leq k are arbitrary numbers (i.e. the chains obtained by dropping the elements of σ\sigma), 0qk0\leq q\leq k.

Definition 3.11.

[st-86] The poset PP with the least element 0^\hat{0} is called a simplicial poset if the subposet [0^,x][\hat{0},x] of PP is a Boolean lattice for any xPx\in P. For any element xx of a simplicial poset PP a length l(x)l(x) of xx is the length of a maximal chain in [0^,x][\hat{0},x]. Here l(0^):=0l(\hat{0}):=0. Define the dimension of a simplicial poset PP to be the number dimP:=dimΔ(P)=maxxPl(x)\dim P:=\dim\Delta(P)=\max_{x\in P}l(x). For a simplicial poset PP let

fi(P):=|{xP|l(x)=i+1}|,f_{i}(P):=|\{x\in P|\ l(x)=i+1\}|,

be the number of elements in PP of length i+1i+1, where i0i\geq 0. In particular, f1(P)=1f_{-1}(P)=1.

Remark 3.12.

The poset SΓopS_{\Gamma}^{op} has the least element Γ\Gamma by the definition and is therefore acyclic. For a torus action with a dense open orbit the poset SMopS_{M}^{op} has the least element and is contractible, too. However, for an arbitrary TT-action on MM the poset SMopS_{M}^{op} is neither acyclic nor simplicial, in general. This is due to the fact that the orbit space QQ is a homological cell complex and the group isomorphism H(Δ(SM))H(Q)H^{*}(\Delta(S_{M}))\cong H^{*}(Q) holds, e.g. see [ma-pa-06, Proposition 5.14]. For instance, it can be checked for the T2T^{2}-action on l(3)\mathcal{F}l(3) (Example 2.6) that the corresponding orbit space is homeomorphic to a sphere QS4Q\cong S^{4}, e.g. see [ay-ma-19].

In the following we need the following well-known Philip Hall’s theorem.

Proposition 3.13.

[ro-64, Proposition 6] Let SS be a simplicial poset of dimension dd. Then the Euler characteristic χ~(Δ(S¯))\tilde{\chi}(\Delta(\overline{S})) of the order complex for S¯:=S0^\overline{S}:=S\setminus\hat{0} in the reduced simplicial homology H~(Δ(S¯))\tilde{H}_{*}(\Delta(\overline{S})) is given by the formula:

χ~(Δ(S¯))=i=1d1(1)ifi(S).\tilde{\chi}(\Delta(\overline{S}))=\sum_{i=-1}^{d-1}(-1)^{i}f_{i}(S).

The computation of Euler characteristic for certain face subposets in SΓS_{\Gamma} for a jj-complete GKM-graph Γ\Gamma is possible (by using Proposition 3.13) due to the following proposition.

Proposition 3.14.

Let Γ\Gamma be a jj-complete GKM-graph for some j1j\geq 1. Then for any jj-face Ξ\Xi of Γ\Gamma the poset (SΓ)Ξop(S_{\Gamma})_{\leq\Xi}^{op} is a simplicial poset of dimension dimΞ\dim\Xi. In particular, for any Ω(SΓ)Ξop\Omega\in(S_{\Gamma})_{\leq\Xi}^{op} one has l(Ω)=jdimΩl(\Omega)=j-\dim\Omega in (SΓ)Ξop(S_{\Gamma})_{\leq\Xi}^{op}, and fi((SΓ)Ξop)f_{i}((S_{\Gamma})_{\leq\Xi}^{op}) is equal to the number of (ji1)(j-i-1)-dimensional faces in (SΓ)Ξ(S_{\Gamma})_{\leq\Xi}.

Proof.

In order to prove the claim of the proposition it is enough to show that for any face Ω\Omega in Ξ\Xi the poset [Ω,Ξ]={ΦSΓ|ΩΦΞ}[\Omega,\Xi]=\{\Phi\in S_{\Gamma}|\ \Omega\subseteq\Phi\subseteq\Xi\} is isomorphic to the poset of faces in ΔjdimΩ\Delta^{j-\dim\Omega}. Let vVΩv\in V_{\Omega}. Since the connection of Γ\Gamma is jj-independent, any face Φ[Ω,Ξ]\Phi\in[\Omega,\Xi] is uniquely determined by the collection C(Φ)C(\Phi) of dimΦdimΩ\dim\Phi-\dim\Omega mutually different elements from starΞvstarΩv\operatorname{star}_{\Xi}v\setminus\operatorname{star}_{\Omega}v, and vice versa. Moreover, for any Φ1,Φ2[Ω,Ξ]\Phi_{1},\Phi_{2}\in[\Omega,\Xi] one has Φ1Φ2\Phi_{1}\subseteq\Phi_{2} iff C(Φ1)C(Φ2)C(\Phi_{1})\subseteq C(\Phi_{2}). This implies the necessary claim. The proof is complete. ∎

4. A periodic GKM-graph and its quotient

In this section we give a detailed construction of the GKM-graph suitable for the proof of Theorem 1 and study some of its properties.

Construction 4.1 (Graph Γ\Gamma^{\prime}).

Let IdR(x)I^{d}_{R}(x) be the edge graph of the cube

𝕀dR(x):={y=(y1,,yd)d||xiyi|R},\mathbb{I}^{d}_{R}(x):=\{y=(y_{1},\dots,y_{d})\in\mathbb{R}^{d}|\ |x_{i}-y_{i}|\leq R\},

with center at x=(x1,,xd)dx=(x_{1},\dots,x_{d})\in\mathbb{R}^{d} and with edges of length 2R2R. For any d1d\geq 1 we define the graph Γ=Γ(d)\Gamma^{\prime}=\Gamma^{\prime}(d) embedded into d\mathbb{R}^{d} as the union of the following graphs:

  1. (i)

    Graph Id1/6(x)I^{d}_{1/6}(x), where xx runs over all points dd\mathbb{Z}^{d}\subset\mathbb{R}^{d} with integral coordinates;

  2. (ii)

    Graph Id1/6(x+12(1,,1))I^{d}_{1/6}(x+\frac{1}{2}\cdot(1,\dots,1)), where xx runs over d\mathbb{Z}^{d};

  3. (iii)

    A diagonal, that is, an edge of the form

    D(x,u):=(x+16i=1d(1)uiei,x+13i=1d(1)uiei),D(x,u):=(x+\frac{1}{6}\cdot\sum_{i=1}^{d}(-1)^{u_{i}}e_{i},\ x+\frac{1}{3}\cdot\sum_{i=1}^{d}(-1)^{u_{i}}e_{i}),

    (and its inverse), where xx runs over d\mathbb{Z}^{d}, u=(u1,,ud)u=(u_{1},\dots,u_{d}) runs over {±1}d\{\pm 1\}^{d}, and e1,,ede_{1},\dots,e_{d} is the standard basis of d\mathbb{R}^{d}.

Notice that the graph Γ\Gamma^{\prime} is a (d+1)(d+1)-regular connected graph with infinite set of vertices. We call a cubical subgraph any subgraph in Γ\Gamma^{\prime} of the form I1/6d(x)I_{1/6}^{d}(x), where xLd:=d(d+12(1,,1))dx\in L_{d}:=\mathbb{Z}^{d}\sqcup(\mathbb{Z}^{d}+\frac{1}{2}(1,\dots,1))\subset\mathbb{R}^{d}. For any vertex vVΓv\in V_{\Gamma^{\prime}} denote by Cube(v)\operatorname{Cube}(v) a unique cubical subgraph in Γ\Gamma^{\prime} such that vVCube(v)v\in V_{\operatorname{Cube}(v)} holds.

Construction 4.2 (Functions εij\varepsilon^{i}_{j}).

For any d1d\geq 1 define the functions εij:VΓ{±1}\varepsilon_{i}^{j}\colon V_{\Gamma^{\prime}}\to\{\ \pm 1\}, where i=1,d+1i=1\dots,d+1 and jj\in\mathbb{N}. For any x=(x1,,xd)dx=(x_{1},\dots,x_{d})\in\mathbb{Z}^{d} and any vertex yy of Id1/6(x)I^{d}_{1/6}(x) let

(4.1) εij(y):=(1)xi2j1;i=1,,d;j.\varepsilon_{i}^{j}(y):=(-1)^{\lfloor\frac{x_{i}}{2^{j-1}}\rfloor};\ i=1,\dots,d;\ j\in\mathbb{N}.

By definition, the function εij\varepsilon_{i}^{j} is then uniquely defined by taking the same values on the vertices of any diagonal of Γ\Gamma^{\prime}, where i=1,,di=1,\dots,d and jj\in\mathbb{N}. Define

(4.2) εd+1j(y):=(1)y12j1,\varepsilon_{d+1}^{j}(y):=(-1)^{\lfloor\frac{y_{1}}{2^{j-1}}\rfloor},

for any vertex y=(y1,,yd)dy=(y_{1},\dots,y_{d})\in\mathbb{Z}^{d} of Γ\Gamma.

Construction 4.3 (Graph Γ\Gamma).

For any d1d\geq 1 and r0r\geq 0 let Γ=Γ(d,r)\Gamma=\Gamma(d,r) be the graph obtained from Γ(d)\Gamma^{\prime}(d) by adding the edges Ej(v):=(v,v)E_{j}(v):=(v,v^{\prime}), where v=v+2j1(1,,1)v^{\prime}=v+2^{j-1}\cdot(1,\dots,1), where jj runs over 1,,r1,\dots,r and vv runs over the subset of elements u=(u1,,ud)VΓu=(u_{1},\dots,u_{d})\in V_{\Gamma^{\prime}} such that (1)iui2j1=1(-1)^{\lfloor\frac{\sum_{i}u_{i}}{2^{j-1}}\rfloor}=1 holds.

Notice that the graph Γ\Gamma is a (d+1+r)(d+1+r)-valent connected graph with infinite set of vertices, and that VΓ=VΓV_{\Gamma}=V_{\Gamma^{\prime}} holds (see Figure LABEL:fig:cross, LABEL:fig:edges).

Construction 4.4 (Axial function α\alpha on Γ\Gamma).

Fix a collection of integers t1,,trt_{1},\dots,t_{r}\in\mathbb{Z}. Let α:EΓd+1\alpha\colon E_{\Gamma}\to\mathbb{Z}^{d+1}, α=α(d,r,t1,,tr)\alpha=\alpha(d,r,t_{1},\dots,t_{r}) be the function taking value α(D(x,u))=ed+1\alpha(D(x,u))=e_{d+1} for any xdx\in\mathbb{Z}^{d} and any u{±1}du\in\{\pm 1\}^{d}. By definition, for any eEId1/6(x)e\in E_{I^{d}_{1/6}(x)} let α(e)\alpha(e) be the inner (outer, respectively) normal of the unit length for the corresponding to ee facet of the cube 𝕀d1/6(x)\mathbb{I}^{d}_{1/6}(x) in d\mathbb{R}^{d} if xdx\in\mathbb{Z}^{d} (if xLddx\in L_{d}\setminus\mathbb{Z}^{d}, respectively). For any diagonal e=D(x,u)e=D(x,u), where xdx\in\mathbb{Z}^{d} and u{±1}du\in\{\pm 1\}^{d}, let α(e)=ed+1\alpha(e)=e_{d+1}, α(e¯)=ed+1\alpha(\overline{e})=-e_{d+1}. In particular, one has α(e){±e1,,±ed+1}\alpha(e)\in\{\pm e_{1},\dots,\pm e_{d+1}\} for any eEΓe\in E_{\Gamma^{\prime}} (see Figure 1). For any vVΓv\in V_{\Gamma} let

(4.3) α(Ej(v))=i=1d+1εij(v)tij1wi(v),\alpha(E_{j}(v))=\sum_{i=1}^{d+1}\varepsilon_{i}^{j}(v)t_{i}^{j-1}w_{i}(v),

for any jj\in\mathbb{N}, where {w1(v),,wd+1(v)}\{w_{1}(v),\dots,w_{d+1}(v)\} are the values of α\alpha on starΓ(v)\operatorname{star}_{\Gamma^{\prime}}(v) denoted in such a way that wi(v)=±eiw_{i}(v)=\pm e_{i} holds for i=1,,d+1i=1,\dots,d+1.

Denote by AxA_{x} the automorphism of the graph Γ(d,r)\Gamma(d,r) induced by the linear operator yx+yy\mapsto x+y in d\mathbb{R}^{d}, where xLdx\in L_{d}; ydy\in\mathbb{R}^{d}. Notice that AxA_{x} is well defined for any xLdx\in L_{d} and that the identities

(4.4) εij(A2j1v(x))=εij(x),\varepsilon_{i}^{j}(A_{2^{j-1}\cdot v}(x))=-\varepsilon_{i}^{j}(x),
(4.5) α(A1/2u(e))=α(e),\alpha(A_{1/2\cdot u}(e))=-\alpha(e),

hold for any xVΓx\in V_{\Gamma}, i=1,,d+1i=1,\dots,d+1; jj\in\mathbb{N}; eEΓe\in E_{\Gamma}; u{±1}du\in\{\pm 1\}^{d} and any vdv\in\mathbb{Z}^{d}.

Our next task is to define the connections \nabla^{\prime}, \nabla on Γ\Gamma^{\prime} and Γ\Gamma compatible with α\alpha^{\prime} and α\alpha by describing the corresponding facets, respectively. We do this by listing all facets in the corresponding graphs in the next two definitions. One can easily check that the facets given below are compatible with α\alpha^{\prime} and α\alpha.

Definition 4.5 (Facets of Γ\Gamma^{\prime}).

For any vVΓv\in V_{\Gamma^{\prime}} let F0(v):=Cube(v)F_{0}(v):=\operatorname{Cube}(v) be the subgraph in Γ\Gamma^{\prime}. Denote by Cubei(v)\operatorname{Cube}_{i}(v) a unique subgraph in Cube(v)\operatorname{Cube}(v) corresponding to the facet of the respective cube 𝕀1/6d(x)\mathbb{I}_{1/6}^{d}(x) with the normal vector ±ei\pm e_{i} such that vCubei(v)v\in\operatorname{Cube}_{i}(v), where i=1,,di=1,\dots,d. Let (u,v)(u,v) be any diagonal of Γ\Gamma^{\prime}. For any i=1,,ni=1,\dots,n let Fdi(v)F^{d}_{i}(v) be the dd-valent subgraph of Γ\Gamma^{\prime} that is the union of subgraphs AaejCubei(u)A_{ae_{j}}\operatorname{Cube}_{i}(u), AaejCubei(v)A_{ae_{j}}\operatorname{Cube}_{i}(v) and AaejeA_{ae_{j}}e, where ee runs over 2d12^{d-1} diagonals of Γ\Gamma^{\prime} incident to Cubei(v)\operatorname{Cube}_{i}(v), jj runs over {1,,d}{i}\{1,\dots,d\}\setminus\{i\} and aa runs over \mathbb{Z}.

Definition 4.6 (Facets of Γ\Gamma).

For any i=0,,di=0,\dots,d let Gi(v)G_{i}(v) be the union of the subgraphs Fi(v)F_{i}(v), Fi(Aq2j1(1,,1)(v))F_{i}(A_{q\cdot 2^{j-1}\cdot(1,\dots,1)}(v)) and edges Ej(x)E_{j}(x) (and their inverses), where jj runs over 1,,r1,\dots,r, qq runs over \mathbb{Z} and xx runs over the union of all vertices of these graphs. For any j=1,,rj=1,\dots,r define the subgraph Gn+j(v)G_{n+j}(v) of Γ\Gamma to be obtained by omitting all edges Ej(v)E_{j}(v) in Γ\Gamma, where vv runs over VΓV_{\Gamma}.

Proposition 4.7.

Let eEΓe\in E_{\Gamma^{\prime}} be any edge such that α(e)=±ei\alpha(e)=\pm e_{i} holds for some i{1,,d+1}i\in\{1,\dots,d+1\}. Then one has

(4.6) εjq(i(e))=εjq(t(e)),q;j=1,,d+1;ji.\varepsilon_{j}^{q}(i(e))=\varepsilon_{j}^{q}(t(e)),\ q\in\mathbb{N};\ j=1,\dots,d+1;\ j\neq i.
Proof.

By the definition, the values of εd+1q\varepsilon_{d+1}^{q} on the vertices of I1/6d(x)I_{1/6}^{d}(x) are equal to each other for any xLdx\in L_{d}. This proves (4.6) for j=d+1j=d+1 (if idi\leq d). By definition, the values of εjq\varepsilon_{j}^{q} are mutually equal on the vertices of the graph I1/6d(x)I_{1/6}^{d}(x), as well as on the vertices of any diagonal emanating from I1/6d(x)I_{1/6}^{d}(x), where j=1,,dj=1,\dots,d; qq\in\mathbb{N} and xdx\in\mathbb{Z}^{d}. Suppose that eEId1/6(x)e\in E_{I^{d}_{1/6}(x)} holds for some xLndx\in L_{n}\setminus\mathbb{Z}^{d}. Notice that idi\leq d holds. Without loss of generality, let e=(i(e),i(e)+1/3ei)e=(i(e),i(e)+1/3e_{i}) and let jdj\leq d. Let (v1,i(e))(v_{1},i(e)), (v2,t(e))(v_{2},t(e)) be both unique diagonals of Γ\Gamma^{\prime} terminating at i(e)i(e) and t(e)t(e), respectively. Notice that AeiCube(v1)=Cube(v2)A_{e_{i}}\operatorname{Cube}(v_{1})=\operatorname{Cube}(v_{2}) holds for the respective subgraphs in Γ\Gamma^{\prime}. Then one uses Construction 4.2 and (4.1) to conduct the following computation

εjq(t(e))=εjq(v2)=εjq(Aei(v1))=εjq(v1)=εjq(i(e)),ji;q.\varepsilon_{j}^{q}(t(e))=\varepsilon_{j}^{q}(v_{2})=\varepsilon_{j}^{q}(A_{e_{i}}(v_{1}))=\varepsilon_{j}^{q}(v_{1})=\varepsilon_{j}^{q}(i(e)),\ j\neq i;\ q\in\mathbb{N}.

The proof is complete. ∎

Lemma 4.8.

The function α\alpha satisfies the rank, opposite and congruence conditions with respect to Γ\Gamma and \nabla (see Definition 2.1).

Proof.

Notice that the rank condition is satisfied for α\alpha by the construction. Let eEΓe\in E_{\Gamma}. Consider the following cases.

1) Let eEΓe\in E_{\Gamma^{\prime}}. The opposite sign condition is easily deduced for α\alpha along the edge ee. In terms of Construction 4.4 of α\alpha, one has α(e)=±eq\alpha(e)=\pm e_{q} for some q=1,,d+1q=1,\dots,d+1. Then one has wi(i(e))=wi(t(e))w_{i}(i(e))=w_{i}(t(e)) for any iqi\neq q and wq(i(e))wq(t(e))0(modeq)w_{q}(i(e))\equiv w_{q}(t(e))\equiv 0\pmod{e_{q}}. Hence, by Proposition 4.7 and by (4.3) one has

(4.7) α(Ej(t(e)))=i=1d+1εij(t(e))tij1wi(t(e))i=1d+1εij(t(e))tij1wi(i(e))i=1d+1εij(i(e))tij1wi(i(e))=α(Ej(i(e)))(modeq),\alpha(E_{j}(t(e)))=\sum_{i=1}^{d+1}\varepsilon_{i}^{j}(t(e))t_{i}^{j-1}w_{i}(t(e))\equiv\sum_{i=1}^{d+1}\varepsilon_{i}^{j}(t(e))t_{i}^{j-1}w_{i}(i(e))\equiv\\ \sum_{i=1}^{d+1}\varepsilon_{i}^{j}(i(e))t_{i}^{j-1}w_{i}(i(e))=\alpha(E_{j}(i(e)))\pmod{e_{q}},

where j=1,,rj=1,\dots,r. Hence, the congruence condition holds for α\alpha along the edge ee.

2) Let eEΓe\notin E_{\Gamma^{\prime}}. Let u=i(e)u=i(e). Then e=Eq(u)e=E_{q}(u) for some q=1,,rq=1,\dots,r. By Construction 4.4, the equality wi(i(e))=wi(t(e))w_{i}(i(e))=w_{i}(t(e)) holds for any i=1,,d+1i=1,\dots,d+1. By (4.4) and (4.3) then one has

α(Eq(i(e))¯)=α(Eq(t(e)))=i=1d+1εiq(t(e))tiq1wi(t(e))=i=1d+1εiq(t(e))tiq1wi(i(e))=i=1d+1εiq(i(e))tiq1wi(i(e))=α(Eq(i(e))).\alpha(\overline{E_{q}(i(e))})=\alpha(E_{q}(t(e)))=\sum_{i=1}^{d+1}\varepsilon_{i}^{q}(t(e))t_{i}^{q-1}w_{i}(t(e))=\\ \sum_{i=1}^{d+1}\varepsilon_{i}^{q}(t(e))t_{i}^{q-1}w_{i}(i(e))=-\sum_{i=1}^{d+1}\varepsilon_{i}^{q}(i(e))t_{i}^{q-1}w_{i}(i(e))=-\alpha(E_{q}(i(e))).

Hence, the opposite sign condition holds for α\alpha along the edge ee. If r=1r=1, then the congruence relations hold for α\alpha along ee. Suppose that r2r\geq 2 holds. Choose any j=1,,rj=1,\dots,r such that jqj\neq q holds. Then by (4.4) one has

εiq(t(e))=εiq(A2q1(1,,1)(i(e)))=εiq((A2min{j,q}1(1,,1))2|jq|(i(e)))=(1)2|jq|εiq(i(e))=εiq(i(e)),\varepsilon_{i}^{q}(t(e))=\varepsilon_{i}^{q}(A_{2^{q-1}\cdot(1,\dots,1)}(i(e)))=\varepsilon_{i}^{q}((A_{2^{\min\{j,q\}-1}\cdot(1,\dots,1)})^{2^{|j-q|}}(i(e)))=\\ (-1)^{2^{|j-q|}}\varepsilon_{i}^{q}(i(e))=\varepsilon_{i}^{q}(i(e)),

for any q=1,,d+1q=1,\dots,d+1. Hence, by (4.3), the computation (4.7) holds in this case. This implies that the congruence condition holds for α\alpha along the edge ee. The proof is complete. ∎

Lemma 4.9.

Let r1r\geq 1. Then there exist integers t1,,trt_{1},\dots,t_{r}\in\mathbb{Z} such that the axial function α(d,r,t1,,tr)\alpha(d,r,t_{1},\dots,t_{r}) is (d+1)(d+1)-independent.

Proof.

For any vertex vVΓv\in V_{\Gamma} the values of the axial function α\alpha on star(v)\operatorname{star}(v) are given by the columns of the following (d+1)×(d+1+r)(d+1)\times(d+1+r)-matrix:

(4.8) ((1)i10(1)i1ε11(v)(1)i1ε1r(v)t1r10(1)id+1(1)id+1εd+11(v)(1)id+1εd+1r(v)td+1r1),\begin{pmatrix}(-1)^{i_{1}}&\cdots&0&(-1)^{i_{1}}\varepsilon_{1}^{1}(v)&\cdots&(-1)^{i_{1}}\varepsilon_{1}^{r}(v)t_{1}^{r-1}\\ \vdots&\ddots&\vdots&\vdots&\ddots&\vdots\\ 0&\cdots&(-1)^{i_{d+1}}&(-1)^{i_{d+1}}\varepsilon_{d+1}^{1}(v)&\cdots&(-1)^{i_{d+1}}\varepsilon_{d+1}^{r}(v)t_{d+1}^{r-1}\end{pmatrix},

where wq(v)=(1)iqeqw_{q}(v)=(-1)^{i_{q}}e_{q} for q=1,,d+1q=1,\dots,d+1 in terms of Construction 4.4, and i1,,id+1i_{1},\dots,i_{d+1} depend on vv. By slightly abusing the notation let M=M(v;j1,,jd+1)M=M(v;\ j_{1},\dots,j_{d+1}) be the (d+1)×(d+1)(d+1)\times(d+1)-minor of the above matrix (4.8) corresponding to the columns with indices 1j1<<jd+1d+1+r1\leq j_{1}<\dots<j_{d+1}\leq d+1+r in (4.8) (from left to right). For any integers 1j1<<jd+1d+1+r1\leq j_{1}<\dots<j_{d+1}\leq d+1+r there exists an integer q{0,,d+1}q\in\{0,\dots,d+1\} such that the inequalities j1,,jqd+1j_{1},\dots,j_{q}\leq d+1 and jq+1,,jd+1>n+1j_{q+1},\dots,j_{d+1}>n+1 hold, where j0:=0j_{0}:=0. If q=n+1q=n+1 then

detM=p=1d+1(1)ip.\det M=\prod_{p=1}^{d+1}(-1)^{i_{p}}.

Let qnq\leq n. The ordering t1<<trt_{1}<\dots<t_{r} of the variables induces the lexicographical ordering on the polynomials from the ring [t1,,tr]\mathbb{Z}[t_{1},\dots,t_{r}]. In this ordering the maximal monomial in detM\det M is equal to

p=1q(1)jps=q+1d+1(1)isεsjs(v)tsjs1.\prod_{p=1}^{q}(-1)^{j_{p}}\cdot\prod_{s=q+1}^{d+1}(-1)^{i_{s}}\varepsilon_{s}^{j_{s}}(v)t_{s}^{j_{s}-1}.

In particular, detM\det M is a nonzero polynomial in t1,,trt_{1},\dots,t_{r}. Hence, the left-hand sides in the system of inequalities detM(v;j1,,jd+1)0\det M(v;\ j_{1},\dots,j_{d+1})\neq 0, where (j1,,jd+1)(j_{1},\dots,j_{d+1}) exhausts all (d+1)(d+1)-subsets of {1,,d+r+1}\{1,\dots,d+r+1\} and vv runs over VΓV_{\Gamma}, includes no zero polynomials. The set of real solutions for this system is the complement to the finite union of subsets of zero measure in r\mathbb{R}^{r}, because α\alpha is periodic (see (4.5)). Therefore, this complement has a rational point with the corresponding coordinates t1,,trt^{\prime}_{1},\dots,t^{\prime}_{r}\in\mathbb{Q}. By multiplying t1,,trt^{\prime}_{1},\dots,t^{\prime}_{r} with the respective least common multiple one obtains t1,,trt_{1},\dots,t_{r}\in\mathbb{Z} such that α(n,r,t1,,tr)\alpha(n,r,t_{1},\dots,t_{r}) is (n+1)(n+1)-independent. This completes the proof. ∎

Remark 4.10.

The values of the axial function α(d,r,t1,,tr)\alpha(d,r,t_{1},\dots,t_{r}) obtained in Lemma 4.9 may not be primitive, in general. However, one can replace each non-primitive value of α(n,r,t1,,tr)\alpha(n,r,t_{1},\dots,t_{r}) with the corresponding primitive vector in d+1\mathbb{Z}^{d+1}. Notice that the axial function obtained during this procedure is (d+1)(d+1)-independent and its values at any star of Γ\Gamma contain a basis of d+1\mathbb{Z}^{d+1}.

e1e_{1}e2e_{2}e2e_{2}e1-e_{1}e2-e_{2}e1e_{1}e1-e_{1}e2-e_{2}e1e_{1}e2e_{2}e2e_{2}e1-e_{1}e2-e_{2}e1e_{1}e1-e_{1}e2-e_{2}e1e_{1}e2e_{2}e2e_{2}e1-e_{1}e2-e_{2}e1e_{1}e1-e_{1}e2-e_{2}e1e_{1}e2e_{2}e2e_{2}e1-e_{1}e2-e_{2}e1e_{1}e1-e_{1}e2-e_{2}e1-e_{1}e2-e_{2}e3-e_{3}e2-e_{2}e1e_{1}e3-e_{3}e3-e_{3}e2e_{2}e1-e_{1}e3-e_{3}e1e_{1}e2e_{2}e3e_{3}e3e_{3}e3e_{3}e3e_{3}
Figure 1. Values of the axial function α\alpha on Γ(2,0)\Gamma(2,0)
Construction 4.11 (GKM-graph Γa\Gamma^{a}).

For any aa\in\mathbb{Z} define an equivalence relation a\sim_{a} on n\mathbb{R}^{n} by putting xayx\sim_{a}y for any x,ynx,y\in\mathbb{R}^{n} such that x=y+ux=y+u for some uadu\in a\cdot\mathbb{Z}^{d}. For any a=b2ra=b\cdot 2^{r}, bb\in\mathbb{Z}, define the graph Γa=Γa(d,r)\Gamma^{a}=\Gamma^{a}(d,r) to be the quotient of Γ\Gamma by a\sim_{a}. Define the axial function αa=αa(d,r,t1,,tr)\alpha^{a}=\alpha^{a}(d,r,t_{1},\dots,t_{r}) and the connection a=a(d,r)\nabla^{a}=\nabla^{a}(d,r) to be induced by α\alpha and \nabla on the quotient graph on Γa\Gamma^{a}, respectively (see (4.5) and Fig. LABEL:fig:conng2).

Example 4.12.

For any d1d\geq 1 the GKM-graph Γ1(d,0)\Gamma^{1}(d,0) is isomorphic to the edge graph of the standard (d+1)(d+1)-dimensional cube 𝕀d+11(0)\mathbb{I}^{d+1}_{1}(0) with the axial function induced by the embedding of 𝕀d+11(0)\mathbb{I}^{d+1}_{1}(0) to d+1\mathbb{R}^{d+1}.

By slightly abusing the notation let []=[]a:dd/a[-]=[-]_{a}\colon\mathbb{R}^{d}\to\mathbb{R}^{d}/\sim_{a} be the quotient map.

Proposition 4.13.

For any a=b2ra=b\cdot 2^{r}, bb\in\mathbb{Z}, the GKM-graph Γa(d,r)\Gamma^{a}(d,r) is a well defined graph with finitely many vertices and edges. Furthermore, it has neither multiple edges nor loops, and has type (d+1+r,d+1)(d+1+r,d+1). The objects [Ax][A_{x}] where xLdx\in L_{d}; [εij][\varepsilon_{i}^{j}] where j=1,,rj=1,\dots,r; [Fni(v)][F^{n}_{i}(v)] are well defined for any bb\in\mathbb{Z}.

Proof.

The quotient d/a\mathbb{R}^{d}/\sim_{a} is obtained by gluing all pairs of opposite facets of the cube 𝕀na/2(0)\mathbb{I}^{n}_{a/2}(0) by respective translations in d\mathbb{R}^{d}. Hence, VΓaV_{\Gamma^{a}} is identified with [VΓ𝕀da/2(0)][V_{\Gamma}\cap\mathbb{I}^{d}_{a/2}(0)]. It follows from (4.4) that [εij][\varepsilon_{i}^{j}] is well defined for any i=1,,d+1i=1,\dots,d+1; j=1,,rj=1,\dots,r. The graph Γa\Gamma^{a} has neither loops nor multiple edges because the integral distance between distinct vertices of any its edge (with respect to d\mathbb{Z}^{d}) is less or equal to 2r12^{r-1} and the integral length of any nonzero element from ada\cdot\mathbb{Z}^{d} is greater or equal to 2r2^{r}. Clearly, the automorphism [Ax][A_{x}] of Γa\Gamma^{a} is well-defined for any xLdx\in L_{d}. We check that for any vVΓv\in V_{\Gamma} the edges [Ei(v)][E_{i}(v)], i=1,,ri=1,\dots,r, are distinct. By applying [Ax[A_{x}] for some xLdx\in L_{d} without loss of generality assume that vVCubeyv\in V_{\operatorname{Cube}{y}}, where y[a2(1,,1),a12(1,,1)]y\in[-\frac{a}{2}(1,\dots,1),\frac{a-1}{2}\cdot(1,\dots,1)] belongs to the big diagonal of the respective cube. It follows from the definition that [Ei(v)][E_{i}(v)], i=1,,ri=1,\dots,r, are distinct for any vVCubeyv\in V_{\operatorname{Cube}{y}} and any y[a2(1,,1),a12(1,,1)]y\in[-\frac{a}{2}(1,\dots,1),\frac{a-1}{2}\cdot(1,\dots,1)]. The proof is complete. ∎