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

Delzant type theorem for torus-equivariantly embedded toric hypersurfaces

Kentaro Yamaguchi Department of Mathematical Sciences, Tokyo Metropolitan University, 1-1 Minami-Ohsawa, Hachioji, Tokyo, 192-0397, Japan [email protected]
Abstract.

In the previous work, we study the closure of a complex subtorus in a toric manifold given by the data of an affine subspace. We call it torus-equivariantly embedded toric manifold when the closure of a complex subtorus is a smooth complex submanifold. In this paper, we clearify the condition for nonsingularity of the closure of a complex subtorus in terms of polytopes. The main result is a generalization of Delzant theorem to the case of torus-equivariantly embedded toric hypersurfaces.

Key words and phrases:
Delzant correspondence, toric Kähler manifold, torus-equivariantly embedding
2020 Mathematics Subject Classification:
Primary 53C40; Secondary 53D20, 14M25

1. Introduction

Symplectic toric manifolds are 2n2n-dimensional symplectic manifolds equipped with the effective Hamiltonian action of an nn-dimensional torus. The image of the moment map for the Hamiltonian torus action is the convex hull of the image of the fixed points of the Hamiltonian action [Ati82, GS82]. Due to the work of Delzant [Del88], there is a one-to-one correspondence between symplectic toric manifolds and certain convex polytopes appeared as the moment polytopes of symplectic toric manifolds. The moment polytopes of symplectic toric manifolds are called Delzant polytopes. Moreover, symplectic toric manifolds canonnically admit a Kähler metric, which is called a Guillemin metric [Gui94a, Gui94b, CDG03]. Using the data of a Delzant polytope, the complements of the toric diviors DD in the corresponding symplectic toric manifold XX can be identified with the complex torus ()n(\mathbb{C}^{\ast})^{n}.

In [Yam24b], we construct the closure C(V)¯\overline{C(V)} of the kk-dimensional complex subtorus C(V)C(V) in the toric divior complement XD()nX\setminus D\cong(\mathbb{C}^{\ast})^{n} from the data of a kk-dimensional affine subspace V𝔱nnV\subset\mathfrak{t}^{n}\cong\mathbb{R}^{n}. The closure C(V)¯\overline{C(V)} can be expressed by the zero locus of polynomials fk+1λ,,fnλf^{\lambda}_{k+1},\ldots,f^{\lambda}_{n} for each vertex λ\lambda of the Delzant polytope, which are determined by the data of the Delzant polytope and the affine subspace. Note that the closure C(V)¯\overline{C(V)} might have singularity at the intersection with the toric diviors DD. We show in [Yam24b, Theorem 4.20] that if the closure C(V)¯\overline{C(V)} is a smooth kk-dimensional complex submanifold in the symplectic toric manifold XX, then the moment polytope for the Hamiltonian kk-dimensional torus action on C(V)¯\overline{C(V)} coincides with the moment polytope for the Hamiltonian kk-dimensional subtorus action on XX. Moreover, the submanifold C(V)¯\overline{C(V)} is a symplectic toric manifold with respect to the kk-dimensional torus action on C(V)¯\overline{C(V)}. In particular, if C(V)¯\overline{C(V)} is a complex submanifold, then the moment polytope for the Hamiltonian kk-dimensional subtorus action on XX is a Delzant polytope. We call the complex submanifold C(V)¯\overline{C(V)} the torus-equivariantly embedded toric manifold.

Even though the moment polytope for the Hamiltonian kk-dimensional subtorus action on XX is a Delzant polytope as a convex polytope for some affine subspace VV, C(V)¯\overline{C(V)} might not be a smooth complex submanifold. For example, let X=P2X=\mathbb{C}P^{2}. For any 11-dimensional affine subspace VV in 𝔱22\mathfrak{t}^{2}\cong\mathbb{R}^{2}, the moment polytope for the Hamiltonian one-dimensional subtorus action associated with VV becomes an interval in (𝔱1)(\mathfrak{t}^{1})^{\ast}\cong\mathbb{R}, which can be seen as the Delzant polytope of P1\mathbb{C}P^{1}. However, there are many affine subspaces VV such that the closure C(V)¯\overline{C(V)} has a singularity (see [Yam24b, Section 5.1]).

1.1. Main Results

In this paper, we show that a Delzant type correspondence for torus-equivariantly embedded toric manifolds of codimension one, i.e. k=n1k=n-1.

Let VV be an (n1)(n-1)-dimensional affine subspace in 𝔱nn\mathfrak{t}^{n}\cong\mathbb{R}^{n} with rational slope, i.e. V¯nn1\overline{V}\cap\mathbb{Z}^{n}\cong\mathbb{Z}^{n-1}, where V¯\overline{V} is the linear part of VV. Then, we may take a primitive \mathbb{Z}-basis p1,,pn1np_{1},\ldots,p_{n-1}\in\mathbb{Z}^{n} of V¯n\overline{V}\cap\mathbb{Z}^{n} and ana\in\mathbb{R}^{n} such that V=p1++pn1+aV=\mathbb{R}p_{1}+\cdots+\mathbb{R}p_{n-1}+a. We define the injective homomorphism iV:Tn1Tni_{V}:T^{n-1}\to T^{n} by

iV(t1,,tn1):=(l=1n1tlpl,e1,,l=1n1tlpl,en),i_{V}(t_{1},\ldots,t_{n-1}):=\left(\prod_{l=1}^{n-1}t_{l}^{\langle p_{l},e_{1}\rangle},\ldots,\prod_{l=1}^{n-1}t_{l}^{\langle p_{l},e_{n}\rangle}\right),

where e1,,ene_{1},\ldots,e_{n} are the standard basis of (𝔱n)n(\mathfrak{t}^{n})^{\ast}\cong\mathbb{R}^{n}. We introduce some notions concerning the pair of a Delzant polytope and an affine subspace.

Definition 1.1.

Let Δ\Delta be a Delzant polytope and V=p1++pn1+aV=\mathbb{R}p_{1}+\cdots+\mathbb{R}p_{n-1}+a an (n1)(n-1)-dimensional affine subspace in 𝔱n\mathfrak{t}^{n} with rational slope. A vertex λ\lambda of the polytope Δ\Delta is a good vertex with respect to the pullback iV:(𝔱n)(𝔱n1)i_{V}^{\ast}:(\mathfrak{t}^{n})^{\ast}\to(\mathfrak{t}^{n-1})^{\ast} if the vertex λ\lambda satisfies the two conditions:

  1. (1)

    iV(λ)i_{V}^{\ast}(\lambda) is a vertex of the convex polytope iV(Δ)i_{V}^{\ast}(\Delta),

  2. (2)

    for the direction vectors v1λ,,vnλv^{\lambda}_{1},\ldots,v^{\lambda}_{n} of the edges from the vertex λ\lambda of the polytope Δ\Delta, the vectors iV(v1λ),,iV(vnλ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n}) in (𝔱n1)(\mathfrak{t}^{n-1})^{\ast} are all nonzero.

Definition 1.2.

Let V=p1++pn1+aV=\mathbb{R}p_{1}+\cdots+\mathbb{R}p_{n-1}+a be an (n1)(n-1)-dimensional affine subspace in 𝔱n\mathfrak{t}^{n} with rational slope. A Delzant polytope Δ\Delta is a good polytope with respect to the map iVi_{V}^{\ast} if any good vertex λ\lambda with respect to the map iVi_{V}^{\ast} satisfies the two conditions:

  • we can choose n1n-1 direction vectors v1λ,,vj1λ,vj+1λ,,vnλv^{\lambda}_{1},\ldots,v^{\lambda}_{j-1},v^{\lambda}_{j+1},\ldots,v^{\lambda}_{n} such that iV(v1λ),,iV(vj1λ),iV(vj+1λ),,iV(vnλ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{j-1}),i_{V}^{\ast}(v^{\lambda}_{j+1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n}) are linearly independent and iV({i=1naiviλai0})={ijbiiV(viλ)bi0}i_{V}^{\ast}(\{\sum_{i=1}^{n}a_{i}v^{\lambda}_{i}\mid a_{i}\geq 0\})=\{\sum_{i\neq j}b_{i}i_{V}^{\ast}(v^{\lambda}_{i})\mid b_{i}\geq 0\},

  • the direction vectors iV(v1λ),,iV(vj1λ),iV(vj+1λ),,iV(vnλ)n1i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{j-1}),i_{V}^{\ast}(v^{\lambda}_{j+1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n})\in\mathbb{Z}^{n-1} from the vertex iV(λ)i_{V}^{\ast}(\lambda) of iV(Δ)i_{V}^{\ast}(\Delta) form a \mathbb{Z}-basis of n1\mathbb{Z}^{n-1}.

Note that the first condition in Definition 1.2 are equivalent to the condition that the polytope iV(Δ)i_{V}^{\ast}(\Delta) is simple.

The closure C(V)¯\overline{C(V)} can be expressed by the zero locus of a polynomial fλf^{\lambda} for each vertex λ\lambda of Δ\Delta. Our first main result is as follows:

Theorem 1.3 (Theorem 3.15).

Let Δ\Delta be a Delzant polytope and V=p1++pn1+aV=\mathbb{R}p_{1}+\cdots+\mathbb{R}p_{n-1}+a an (n1)(n-1)-dimensional affine subspace in 𝔱n\mathfrak{t}^{n} with rational slope. If the Delzant polytope Δ\Delta is good with respect to the map iVi_{V}^{\ast}, then the rank of the Jacobian matrix DfλDf^{\lambda} is equal to one at any point of the zero locus of fλf^{\lambda} for any vertex λ\lambda of the Delzant polytope Δ\Delta. In particular, C(V)¯\overline{C(V)} is a smooth complex hypersurface in the toric manifold XX.

Since complex hypersurfaces C(V)¯\overline{C(V)} are symplectic toric manifolds with respect to the Hamiltonian Tn1T^{n-1}-action, we obtain the following:

Corollary 1.4.

If the Delzant polytope Δ\Delta of a toric manifold XX is good with respect to the map iVi_{V}^{\ast}, then the convex polytope iV(Δ)i_{V}^{\ast}(\Delta) is a Delzant polytope of the complex hypersurface C(V)¯\overline{C(V)} in XX.

We also show the converse of Theorem 1.3.

Theorem 1.5 (Theorem 4.13).

Let V=p1++pn1+aV=\mathbb{R}p_{1}+\cdots+\mathbb{R}p_{n-1}+a be an (n1)(n-1)-dimensional affine subspace in 𝔱n\mathfrak{t}^{n} with rational slope. If the rank of the Jacobian matrix DfλDf^{\lambda} is equal to one at any point of the zero locus of fλf^{\lambda} for any vertex λ\lambda of the Delzant polytope Δ\Delta of the toric manifold XX, then the Delzant polytope Δ\Delta is good with respect to the map iVi_{V}^{\ast}.

From Theorem 1.3 and Theorem 1.5, we obtain the following:

Corollary 1.6.

Let Δ\Delta be a Delzant polytope of a symplectic toric manifold XX. Then, C(V)¯\overline{C(V)} is a torus-equivariantly embedded toric hypersurface in XX if and only if Δ\Delta is a good Delzant polytope with respect to the map iVi_{V}^{\ast}.

Corollary 1.6 characterizes the condition when C(V)¯\overline{C(V)} is a smooth complex hypersurface in terms of the combinatorics of Delzant polytopes.

1.2. Sketch of the Proofs

We explain the sketch of the proof of Theorem 1.3. First, in Section 3.1 we show that if the vertex λ\lambda of the Delzant polytope Δ\Delta does not satisfy the condition (2) in Definition 1.1, then the rank of the Jacobian matrix DfλDf^{\lambda} is equal to one. Next, in Section 3.2 we show that if the vertex λ\lambda satisfies the condition (2) but does not satisfy the condition (1) in Definition 1.1, then the rank of the Jacobian matrix DfλDf^{\lambda} is equal to one. Finally, in Section 3.3 we consider the case when the vertex λ\lambda is good with respect to the map iVi_{V}^{\ast}. The detailed proof of Theorem 1.3 is given in Section 3.4.

The proof of Theorem 1.5 is based on the observation of the conditions of the sets λ+,λ\mathcal{I}^{+}_{\lambda},\mathcal{I}^{-}_{\lambda}, which are defined in Definition 2.3.

1.3. Outline

This paper is organized as follows. In Section 2, we prepare for something about torus-equivariantly embedded toric hypersurfaces. In Section 3, we show Theorem 1.3. In Section 4, we show Theorem 1.5.

Acknowlegements

The author is grateful to the advisor, Manabu Akaho, for a lot of encouragements and supports. The author would also like to thank Yuichi Kuno and Yasuhito Nakajima for helpful discussions. This work was supported by JST SPRING, Grant Number JPMJSP2156.

2. Reveiw of Torus-equivariantly Embedded Toric Manifolds

In this section, we review the settings in [Yam24b]. We also show some properties of the rank of the Jacobian matrix DfλDf^{\lambda} used in this paper.

2.1. Symplectic Toric Manifolds

In [Del88], symplectic toric manifolds are completely classified by the certain convex polytopes, known as Delzant polytopes:

Definition 2.1.

A convex polytope Δ\Delta in (𝔱n)n(\mathfrak{t}^{n})^{\ast}\cong\mathbb{R}^{n} is Delzant if Δ\Delta satisfies the following three properties:

  • simple; each vertex has nn edges,

  • rational; the direction vectors v1λ,,vnλv^{\lambda}_{1},\ldots,v^{\lambda}_{n} of the edges from any vertex λΛ\lambda\in\Lambda can be chosen as integral vectors,

  • smooth; the vectors v1λ,,vnλnv^{\lambda}_{1},\ldots,v^{\lambda}_{n}\in\mathbb{Z}^{n} chosen as above form a \mathbb{Z}-basis of n\mathbb{Z}^{n}.

Here, Λ\Lambda denotes the set of the vertices of Δ\Delta.

Some literatures (for example, [CDG03]) define Delzant polytopes in terms of the inward pointing normal vectors u1λ,,unλu^{\lambda}_{1},\ldots,u^{\lambda}_{n} to the nn facets sharing each vertex instead of the direction vectors from each vertex. In fact, these two ways to define Delzant polytopes are equivalent because of the following equation:

[u1λunλ]t[v1λvnλ]=En{{}^{t}[u^{\lambda}_{1}\;\cdots\;u^{\lambda}_{n}]}[v^{\lambda}_{1}\;\cdots\;v^{\lambda}_{n}]=E_{n}

holds for any vertex λ\lambda of Δ\Delta (see [Yam24b, Lemma 3.10] for example).

Due to the Delzant construction, we obtain the corresponding symplectic toric manifold XX from the data of a given Delzant polytope Δ\Delta. Here we give the expression of a system of the inhomogeneous coordinate charts on the corresponding symplectic toric manifold XX from the data of the Delzant polytope Δ\Delta. We define two matrices Qλ(=[Qijλ])Q^{\lambda}(=[Q^{\lambda}_{ij}]) and Dλσ(=[dijλσ])D^{\lambda\sigma}(=[d^{\lambda\sigma}_{ij}]) by

Qλ=[v1λvnλ],Dλσ=(Qλ)1QσQ^{\lambda}=\left[v^{\lambda}_{1}\;\cdots\;v^{\lambda}_{n}\right],\;D^{\lambda\sigma}=(Q^{\lambda})^{-1}Q^{\sigma}

for any vertices λ,σΛ\lambda,\sigma\in\Lambda.

Lemma 2.2.

From the data of a given Delzant polytope Δ\Delta, we can construct an open covering {Uλ}λΛ\{U_{\lambda}\}_{\lambda\in\Lambda} of the corresponding toric manifold XX and a set of maps {φλ:Uλn}λΛ\{\varphi_{\lambda}:U_{\lambda}\to\mathbb{C}^{n}\}_{\lambda\in\Lambda} such that

φσφλ1(zλ)=(j=1n(zjλ)dj1λσ,,j=1n(zjλ)djnλσ)\varphi_{\sigma}\circ\varphi_{\lambda}^{-1}(z^{\lambda})=\left(\prod_{j=1}^{n}(z^{\lambda}_{j})^{d^{\lambda\sigma}_{j1}},\ldots,\prod_{j=1}^{n}(z^{\lambda}_{j})^{d^{\lambda\sigma}_{jn}}\right)

for any vertices λ,σΛ\lambda,\sigma\in\Lambda such that UλUσU_{\lambda}\cap U_{\sigma}\neq\emptyset.

The detailed construction in terms of the notation used here is given in [Yam24b, Section 2.1 and Section 3.2]. Moreover, the complement of toric diviors in a 2n2n-dimensional symplectic toric manifold can be identified with a complex nn-dimensional torus.

2.2. Torus-equivariantly Embedded Toric Hypersurfaces

Let p1,,pn1np_{1},\ldots,p_{n-1}\in\mathbb{Z}^{n} be primitive vectors which are linearly independent, and ana\in\mathbb{R}^{n}. We consider the (n1)(n-1)-dimensional affine subspace V=p1++pn1+aV=\mathbb{R}p_{1}+\cdots+\mathbb{R}p_{n-1}+a in 𝔱nn\mathfrak{t}^{n}\cong\mathbb{R}^{n}. Write V¯\overline{V} as the linear part of VV. Assume that VV has a rational slope, i.e. we may assume that p1,,pn1np_{1},\ldots,p_{n-1}\in\mathbb{Z}^{n} form a \mathbb{Z}-basis of V¯nn1\overline{V}\cap\mathbb{Z}^{n}\cong\mathbb{Z}^{n-1}. Let qnq\in\mathbb{Z}^{n} be a primitive basis of the orthogonal subspace to V¯\overline{V} in 𝔱nn\mathfrak{t}^{n}\cong\mathbb{R}^{n}.

Through the log-affine coordinate (n)𝔱n×Tn(\mathbb{C}^{n})^{\ast}\cong\mathfrak{t}^{n}\times T^{n}, we define the complex subtorus C(V)V×Tn1C(V)\cong V\times T^{n-1} (see [Yam24b, Proposition 4.2] for detail). Recall that the complement of toric diviors in a toric manifold XX can be identified with a complex nn-dimensional torus ()n(\mathbb{C}^{\ast})^{n}. The closure C(V)¯\overline{C(V)} of C(V)C(V) in the toric manifold XX is expressed as follows:

Definition 2.3.

The closure C(V)¯\overline{C(V)} of C(V)C(V) in the toric manifold XX is defined as C(V)¯=λΛφλ1(Cλ(V)¯)\overline{C(V)}=\bigcup_{\lambda\in\Lambda}\varphi_{\lambda}^{-1}(\overline{C_{\lambda}(V)}), where Cλ(V)¯φλ(Uλ)n\overline{C_{\lambda}(V)}\subset\varphi_{\lambda}(U_{\lambda})\cong\mathbb{C}^{n} is a zero locus of fλf^{\lambda}. Here, fλf^{\lambda} are defined by

fλ(zλ)=jλ+λ0(zjλ)ujλ,qea,qjλλ0(zjλ)ujλ,q,f^{\lambda}(z^{\lambda})=\prod_{j\in\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{\langle u^{\lambda}_{j},q\rangle}-e^{\langle a,q\rangle}\prod_{j\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{-\langle u^{\lambda}_{j},q\rangle},

where

λ+={i{1,,n}uiλ,q0},\displaystyle\mathcal{I}^{+}_{\lambda}=\{i\in\{1,\ldots,n\}\mid\langle u^{\lambda}_{i},q\rangle\geq 0\},
λ={i{1,,n}uiλ,q0},\displaystyle\mathcal{I}^{-}_{\lambda}=\{i\in\{1,\ldots,n\}\mid\langle u^{\lambda}_{i},q\rangle\leq 0\},
λ0={i{1,,n}uiλ,q=0}.\displaystyle\mathcal{I}^{0}_{\lambda}=\{i\in\{1,\ldots,n\}\mid\langle u^{\lambda}_{i},q\rangle=0\}.

Note that if λ+λ0=\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset, then jλ+λ0(zjλ)ujλ,q=1\prod_{j\in\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{\langle u^{\lambda}_{j},q\rangle}=1. Similarly, if λλ0=\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset, then jλλ0(zjλ)ujλ,q=1\prod_{j\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{-\langle u^{\lambda}_{j},q\rangle}=1.

Remark 2.4.

The rank of the Jacobian matrix DfλDf^{\lambda} of fλf^{\lambda} might be equal to zero at some points in the toric diviors. In particular, C(V)¯\overline{C(V)} might have a singularity at some points in the toric diviors.

Proposition 2.5.

If C(V)¯\overline{C(V)} is a complex submanifold in the toric manifold XX, then C(V)¯\overline{C(V)} is toric with respect to the Tn1T^{n-1}-action on C(V)¯\overline{C(V)}.

Proof.

We show that the Tn1T^{n-1}-action on C(V)¯\overline{C(V)} is effective, Hamiltonian. Define the injective homomorphism iV:Tn1Tni_{V}:T^{n-1}\to T^{n} by

iV(t1,,tn1)=(l=1n1tlpl,e1,,l=1n1tlpl,en).i_{V}(t_{1},\ldots,t_{n-1})=\left(\prod_{l=1}^{n-1}t_{l}^{\langle p_{l},e_{1}\rangle},\ldots,\prod_{l=1}^{n-1}t_{l}^{\langle p_{l},e_{n}\rangle}\right).

Since the toric divior complement XDX\setminus D can be identified with a complex torus ()n(\mathbb{C}^{\ast})^{n}, we obtain XD𝔱n×TnX\setminus D\cong\mathfrak{t}^{n}\times T^{n}. Since C(V)()n1C(V)\cong(\mathbb{C}^{\ast})^{n-1} is a complex subtorus in ()nXD(\mathbb{C}^{\ast})^{n}\cong X\setminus D, we obtain the identification C(V)V×iV(Tn1)𝔱n×TnXDC(V)\cong V\times i_{V}(T^{n-1})\subset\mathfrak{t}^{n}\times T^{n}\cong X\setminus D. Since we defined the Tn1T^{n-1}-action on C(V)¯\overline{C(V)} through the map iV:Tn1Tni_{V}:T^{n-1}\to T^{n} and the TnT^{n}-action on XX (see [Yam24b]), we see that the Tn1T^{n-1}-action on C(V)¯\overline{C(V)} is effective.

If C(V)¯\overline{C(V)} is a complex submanifold in XX, then there exists an embedding i:C(V)¯Xi:\overline{C(V)}\to X. Thus, a symplectic form on C(V)¯\overline{C(V)} is iωi^{\ast}\omega, where ω\omega is a symplectic form on XX. Let Rt:C(V)¯C(V)¯R_{t}:\overline{C(V)}\to\overline{C(V)} be a map defined by

Rt(x)=xt,R_{t}(x)=x\cdot t,

where \cdot is the Tn1T^{n-1}-action on C(V)¯\overline{C(V)}. By the definition of the Tn1T^{n-1}-action on C(V)¯\overline{C(V)}, we obtain

Rt(iω)=iRtω=iω,R_{t}^{\ast}(i^{\ast}\omega)=i^{\ast}R_{t}^{\ast}\omega=i^{\ast}\omega,

for any tTn1t\in T^{n-1}. Thus, the Tn1T^{n-1}-action on C(V)¯\overline{C(V)} preserves the symplectic form iωi^{\ast}\omega on C(V)¯\overline{C(V)}. We show that the map μ¯:=iVμi:C(V)¯(𝔱n1)\overline{\mu}:=i_{V}^{\ast}\circ\mu\circ i:\overline{C(V)}\to(\mathfrak{t}^{n-1})^{\ast} is the moment map for the Tn1T^{n-1}-action on C(V)¯\overline{C(V)}, where μ:X(𝔱n)\mu:X\to(\mathfrak{t}^{n})^{\ast} is the moment map for the TnT^{n}-action on XX. By the definition of the Tn1T^{n-1}-action on C(V)¯\overline{C(V)}, the map μ¯\overline{\mu} is equivariant with respect to the Tn1T^{n-1}-action on C(V)¯\overline{C(V)}. For any point pC(V)¯p\in\overline{C(V)} and X𝔱n1X\in\mathfrak{t}^{n-1}, we have

μ¯X(p)\displaystyle\overline{\mu}_{X}(p) =iVμ(i(p)),X\displaystyle=\langle i_{V}^{\ast}\mu(i(p)),X\rangle
=μ(i(p)),(iV)X\displaystyle=\langle\mu(i(p)),(i_{V})_{\ast}X\rangle
=μ(i(p)),X\displaystyle=\langle\mu(i(p)),X\rangle
=(μXi)(p).\displaystyle=(\mu_{X}\circ i)(p).

From this calculation, we obtain

dμ¯X\displaystyle d\overline{\mu}_{X} =d(μXi)\displaystyle=d(\mu_{X}\circ i)
=diμX\displaystyle=di^{\ast}\mu_{X}
=idμX\displaystyle=i^{\ast}d\mu_{X}
=iω(X,).\displaystyle=i^{\ast}\omega(X^{\sharp},-).

Therefore, the map μ¯\overline{\mu} is the moment map for the Tn1T^{n-1}-action on C(V)¯\overline{C(V)} equipped with the symplectic form iωi^{\ast}\omega. ∎

Note that the above discussion is valid for any codimension of C(V)¯\overline{C(V)}.

The vectors q,v1λ,,vnλ,u1λ,,unλq,v^{\lambda}_{1},\ldots,v^{\lambda}_{n},u^{\lambda}_{1},\ldots,u^{\lambda}_{n} used here have the following relation:

Lemma 2.6 ([Yam24b, Lemma 4.14]).
j=1nujλ,qiV(vjλ)=0\sum_{j=1}^{n}\langle u^{\lambda}_{j},q\rangle i_{V}^{\ast}(v^{\lambda}_{j})=0

holds.

In this case, the rank of the Jacobian matrix DfλDf^{\lambda} is independent of the choice of ana\in\mathbb{R}^{n} of the affine subspace V=p1++pn1+aV=\mathbb{R}p_{1}+\cdots+\mathbb{R}p_{n-1}+a [Yam24a, Proposition 2.12], i.e. we may assume that a=0a=0.

2.3. Properties of the Rank of the Jacobian Matrices

In this section, we prepare for the proof of the main results.

Lemma 2.7.

Let λ\lambda be a vertex of Δ\Delta.

  • If λ+λ0=\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset, then Cλ(V)¯{zjλ=0}=\overline{C_{\lambda}(V)}\cap\{z^{\lambda}_{j}=0\}=\emptyset for any jλλ0j\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}.

  • If λλ0=\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset, then Cλ(V)¯{zjλ=0}=\overline{C_{\lambda}(V)}\cap\{z^{\lambda}_{j}=0\}=\emptyset for any jλ+λ0j\in\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}.

Proof.

We assume that λ+λ0=\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset. Then, since the defining equation fλ(zλ)=0f^{\lambda}(z^{\lambda})=0 for Cλ(V)¯\overline{C_{\lambda}(V)} is

fλ(zλ)=1jλλ0(zjλ)ujλ,q,f^{\lambda}(z^{\lambda})=1-\prod_{j\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{-\langle u^{\lambda}_{j},q\rangle},

we obtain that Cλ(V)¯{zjλ=0}=\overline{C_{\lambda}(V)}\cap\{z^{\lambda}_{j}=0\}=\emptyset for any jλλ0j\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}.

Similarly, if we assume that λλ0=\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset, then we obtain Cλ(V)¯{zjλ=0}=\overline{C_{\lambda}(V)}\cap\{z^{\lambda}_{j}=0\}=\emptyset for any jλ+λ0j\in\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}. ∎

Lemma 2.8.

If λ+λ0=\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset (or λλ0=\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset), then Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular.

Proof.

Assume that λ+λ0=\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset. From the proof of Lemma 2.7, the defining equation fλ(zλ)=0f^{\lambda}(z^{\lambda})=0 for Cλ(V)¯\overline{C_{\lambda}(V)} is

fλ(zλ)=1jλλ0(zjλ)ujλ,q.f^{\lambda}(z^{\lambda})=1-\prod_{j\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{-\langle u^{\lambda}_{j},q\rangle}.

Since we obtain

fλziλ={uiλ,qjλλ0(zjλ)ujλ,qδijif iλλ0,0otherwise,\frac{\partial f^{\lambda}}{\partial z^{\lambda}_{i}}=\begin{cases*}\langle u^{\lambda}_{i},q\rangle\prod_{j\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{-\langle u^{\lambda}_{j},q\rangle-\delta_{ij}}&if $i\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}$,\\ 0&otherwise,\end{cases*}

the points where the rank of the Jacobian matrix DfλDf^{\lambda} of fλf^{\lambda} is equal to zero should be in Cλ(V)¯(jλλ0{zjλ=0})\overline{C_{\lambda}(V)}\cap(\bigcup_{j\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}\{z^{\lambda}_{j}=0\}). However, since Cλ(V)¯(jλλ0{zjλ=0})=\overline{C_{\lambda}(V)}\cap(\bigcup_{j\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}\{z^{\lambda}_{j}=0\})=\emptyset from Lemma 2.7, the rank of DfλDf^{\lambda} is equal to one, i.e. Cλ(V)¯\overline{C_{\lambda}(V)} does not have a singular point.

Similarly, we assume that λλ0=\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset, then we obtain that the rank of DfλDf^{\lambda} is equal to one. ∎

Lemma 2.9.

Assume that

λ+={1,,n1},λλ0={n}\mathcal{I}^{+}_{\lambda}=\{1,\ldots,n-1\},\;\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\{n\}

or

λ+λ0={n},λ={1,,n1}.\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\{n\},\;\mathcal{I}^{-}_{\lambda}=\{1,\ldots,n-1\}.

If unλ,q=±1\langle u^{\lambda}_{n},q\rangle=\pm 1, then Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular.

Proof.

Assume that λ+={1,,n1},λλ0={n}\mathcal{I}^{+}_{\lambda}=\{1,\ldots,n-1\},\;\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\{n\}. From the defining equation for Cλ(V)¯\overline{C_{\lambda}(V)}, we obtain that

fλziλ={uiλ,qjλ+λ0(zjλ)ujλ,qδijif i=1,,n1,unλ,q(znλ)unλ,q1if i=n.\frac{\partial f^{\lambda}}{\partial z^{\lambda}_{i}}=\begin{cases*}\langle u^{\lambda}_{i},q\rangle\prod_{j\in\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{\langle u^{\lambda}_{j},q\rangle-\delta_{ij}}&if $i=1,\ldots,n-1$,\\ \langle u^{\lambda}_{n},q\rangle(z^{\lambda}_{n})^{-\langle u^{\lambda}_{n},q\rangle-1}&if $i=n$.\end{cases*}

If unλ,q=1\langle u^{\lambda}_{n},q\rangle=-1, then

fλznλ=1.\frac{\partial f^{\lambda}}{\partial z^{\lambda}_{n}}=-1.

Therefore, the rank of the Jacobian matrix DfλDf^{\lambda} is equal to one, i.e. Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular. ∎

3. Delzant Polytopes of Torus-equivariantly Embedded Toric Hypersurfaces

In this section, we show the first main result (Theorem 3.15).

Since iV:(𝔱n)(𝔱n1)i_{V}^{\ast}:(\mathfrak{t}^{n})^{\ast}\to(\mathfrak{t}^{n-1})^{\ast} is surjective, we have the following:

Lemma 3.1.

Let v1,,vnn(𝔱n)v_{1},\ldots,v_{n}\in\mathbb{R}^{n}\cong(\mathfrak{t}^{n})^{\ast} be linearly independent. There exist j1,,jn1{1,,n}j_{1},\ldots,j_{n-1}\in\{1,\ldots,n\} such that iV(vj1),,iV(vjn1)(𝔱n1)n1i_{V}^{\ast}(v_{j_{1}}),\ldots,i_{V}^{\ast}(v_{j_{n-1}})\in(\mathfrak{t}^{n-1})^{\ast}\cong\mathbb{R}^{n-1} are linearly independent.

Hereafter, we assume that j1=1,,jn1=n1j_{1}=1,\ldots,j_{n-1}=n-1 otherwise specified.

3.1. Case of 𝒥λ\mathcal{J}_{\lambda}\neq\emptyset

Let λ\lambda be a vertex of the Delzant polytope Δ\Delta. Define 𝒥λ={ip1,viλ==pn1,viλ=0}\mathcal{J}_{\lambda}=\{i\mid\langle p_{1},v^{\lambda}_{i}\rangle=\cdots=\langle p_{n-1},v^{\lambda}_{i}\rangle=0\}. In this section, we show that if the vertex λ\lambda of Δ\Delta does not satisfy the condition (2) in Definition 1.1, then Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular.

Lemma 3.2.

The vertex λ\lambda of Δ\Delta does not satisfy the condition (2) in Definition 1.1 if and only if 𝒥λ\mathcal{J}_{\lambda}\neq\emptyset.

Proof.

If the vertex λ\lambda of Δ\Delta does not satisfy the condition (2) in Definition 1.1, then there exists a direction vector vjλv^{\lambda}_{j} such that the vector iV(vjλ)(𝔱n1)i_{V}^{\ast}(v^{\lambda}_{j})\in(\mathfrak{t}^{n-1})^{\ast} is zero. Note that the pullback iV:(𝔱n)(𝔱n1)i_{V}^{\ast}:(\mathfrak{t}^{n})^{\ast}\to(\mathfrak{t}^{n-1})^{\ast} is given by

iV(ξ)=(p1,ξ,,pn1,ξ).i_{V}^{\ast}(\xi)=\left(\langle p_{1},\xi\rangle,\ldots,\langle p_{n-1},\xi\rangle\right).

Since iV(vjλ)=(0,,0)i_{V}^{\ast}(v^{\lambda}_{j})=(0,\ldots,0), we obtain p1,vjλ=0,,pn1,vjλ=0\langle p_{1},v^{\lambda}_{j}\rangle=0,\ldots,\langle p_{n-1},v^{\lambda}_{j}\rangle=0, i.e. j𝒥λj\in\mathcal{J}_{\lambda}. In particular, 𝒥λ\mathcal{J}_{\lambda}\neq\emptyset.

If 𝒥λ\mathcal{J}_{\lambda}\neq\emptyset, then there exists an element j0𝒥λj_{0}\in\mathcal{J}_{\lambda}. By the definition of the set 𝒥λ\mathcal{J}_{\lambda}, we have p1,vj0λ=0,,pn1,vj0λ=0\langle p_{1},v^{\lambda}_{j_{0}}\rangle=0,\ldots,\langle p_{n-1},v^{\lambda}_{j_{0}}\rangle=0, i.e. the vector iV(vj0λ)i_{V}^{\ast}(v^{\lambda}_{j_{0}}) is zero. ∎

In particular, 𝒥λ\mathcal{J}_{\lambda}\neq\emptyset is equivalent to |𝒥λ|=1|\mathcal{J}_{\lambda}|=1.

Lemma 3.3.

If 𝒥λ\mathcal{J}_{\lambda}\neq\emptyset, then λ+λ0=\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset or λλ0=\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset.

Proof.

Let j0𝒥λj_{0}\in\mathcal{J}_{\lambda}. By the definition of 𝒥λ\mathcal{J}_{\lambda}, we obtain vj0λV=qv^{\lambda}_{j_{0}}\in V^{\perp}=\mathbb{R}q. Since the vectors vj0λ,qv^{\lambda}_{j_{0}},q are nonzero, there exists b{0}b\in\mathbb{R}\setminus\{0\} such that q=bvj0λq=bv^{\lambda}_{j_{0}}. Since uiλ,vjλ=δij\langle u^{\lambda}_{i},v^{\lambda}_{j}\rangle=\delta_{ij} (see [Yam24b, Lemma 3.10]), we have

uiλ,q=uiλ,bvj0λ=buiλ,vj0λ=bδij0\langle u^{\lambda}_{i},q\rangle=\langle u^{\lambda}_{i},bv^{\lambda}_{j_{0}}\rangle=b\langle u^{\lambda}_{i},v^{\lambda}_{j_{0}}\rangle=b\delta_{ij_{0}}

for any i=1,,ni=1,\ldots,n. If b>0b>0, then iλ+i\in\mathcal{I}^{+}_{\lambda} for any ii, i.e. λλ0=\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset. If b<0b<0, then iλi\in\mathcal{I}^{-}_{\lambda} for any ii, i.e. λ+λ0=\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset. ∎

From Lemma 3.3 and Lemma 2.8, we obtain the following:

Corollary 3.4.

If 𝒥λ\mathcal{J}_{\lambda}\neq\emptyset, then Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular.

For a vertex λ\lambda of Δ\Delta, we define the cone 𝒞λ:={i=1naiviλai0}\mathcal{C}_{\lambda}:=\{\sum_{i=1}^{n}a_{i}v^{\lambda}_{i}\mid a_{i}\geq 0\}.

Proposition 3.5.

If 𝒥λ\mathcal{J}_{\lambda}\neq\emptyset, then iV(λ)i_{V}^{\ast}(\lambda) is a vertex of the convex polytope iV(Δ)i_{V}^{\ast}(\Delta).

Proof.

From Lemma 3.1, we may assume that the vectors iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) are linearly independent. Then, n𝒥λn\in\mathcal{J}_{\lambda}, i.e. iV(vnλ)=0(𝔱n1)i_{V}^{\ast}(v^{\lambda}_{n})=0\in(\mathfrak{t}^{n-1})^{\ast}.

If there does not exist any nontrivial subspace in the cone iV(𝒞λ)i_{V}^{\ast}(\mathcal{C}_{\lambda}), then iV(λ)i_{V}^{\ast}(\lambda) is a vertex in the polytope iV(Δ)i_{V}^{\ast}(\Delta). Let WVλW^{\lambda}_{V} be a subspace contained in the cone iV(𝒞λ)i_{V}^{\ast}(\mathcal{C}_{\lambda}). For any element xWVλiV(𝒞λ)x\in W^{\lambda}_{V}\subset i_{V}^{\ast}(\mathcal{C}_{\lambda}), there exist r1,,rn0r_{1},\ldots,r_{n}\geq 0 such that

x=j=1nrjiV(vjλ)=j=1n1rjiV(vjλ).x=\sum_{j=1}^{n}r_{j}i_{V}^{\ast}(v^{\lambda}_{j})=\sum_{j=1}^{n-1}r_{j}i_{V}^{\ast}(v^{\lambda}_{j}).

Since WVλW^{\lambda}_{V} is a linear space, (x)WVλ(-x)\in W^{\lambda}_{V}. There exist s1,,sn0s_{1},\ldots,s_{n}\geq 0 such that

x=j=1nsjiV(vjλ)=j=1n1sjiV(vjλ).-x=\sum_{j=1}^{n}s_{j}i_{V}^{\ast}(v^{\lambda}_{j})=\sum_{j=1}^{n-1}s_{j}i_{V}^{\ast}(v^{\lambda}_{j}).

Let Rj:=rj+sjR_{j}:=r_{j}+s_{j} for j=1,,nj=1,\ldots,n. Since x+(x)=0x+(-x)=0, we obtain

0=j=1n1RjiV(vjλ).0=\sum_{j=1}^{n-1}R_{j}i_{V}^{\ast}(v^{\lambda}_{j}).

Since we assume that iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) are linearly independent, we obtain that R1==Rn1=0R_{1}=\cdots=R_{n-1}=0. Since r1,,rn1,s1,,sn10r_{1},\ldots,r_{n-1},s_{1},\ldots,s_{n-1}\geq 0, we obtain r1==rn1=s1==sn1=0r_{1}=\cdots=r_{n-1}=s_{1}=\cdots=s_{n-1}=0, which means that x=0x=0. Therefore, we obtain WVλ={0}W^{\lambda}_{V}=\{0\}, i.e. there is no nontrivial subspace in the cone iV(𝒞λ)i_{V}^{\ast}(\mathcal{C}_{\lambda}). The point iV(λ)i_{V}^{\ast}(\lambda) is a vertex of the convex polytope iV(Δ)i_{V}^{\ast}(\Delta). ∎

This proposition tells us that if the vertex λ\lambda of Δ\Delta does not satisfy the condition (2) in Definition 1.1, then the vertex λ\lambda satisfies the condition (1).

Proposition 3.6.

If 𝒥λ\mathcal{J}_{\lambda}\neq\emptyset, then the vectors iV(vj1λ),,iV(vjn1λ)i_{V}^{\ast}(v^{\lambda}_{j_{1}}),\ldots,i_{V}^{\ast}(v^{\lambda}_{j_{n-1}}) for j1,,jn1𝒥λj_{1},\ldots,j_{n-1}\notin\mathcal{J}_{\lambda} are the direction vectors from the vertex iV(λ)i_{V}^{\ast}(\lambda) of the polytope iV(Δ)i_{V}^{\ast}(\Delta).

Proof.

Let j𝒥λj\in\mathcal{J}_{\lambda}, i.e. iV(vjλ)=0i_{V}^{\ast}(v^{\lambda}_{j})=0. From Lemma 3.1, the (n1)(n-1) vectors iV(v1λ),,iV(vj1λ),iV(vj+1λ),,iV(vnλ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{j-1}),i_{V}^{\ast}(v^{\lambda}_{j+1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n}) are linearly independent.

For the rest of the proof, we show that

iV({i=1naiviλai0})={ijbiiV(viλ)bi0}.i_{V}^{\ast}\left(\left\{\sum_{i=1}^{n}a_{i}v^{\lambda}_{i}\mid a_{i}\geq 0\right\}\right)=\left\{\sum_{i\neq j}b_{i}i_{V}^{\ast}(v^{\lambda}_{i})\mid b_{i}\geq 0\right\}.

Since it is clear that the right hand side is a subset of the left hand side, we show the other inclusion. For any a1,,an0a_{1},\ldots,a_{n}\geq 0, we obtain

iV(i=1naiviλ)=i=1naiiV(viλ)=ijaiiV(viλ).i_{V}^{\ast}\left(\sum_{i=1}^{n}a_{i}v^{\lambda}_{i}\right)=\sum_{i=1}^{n}a_{i}i_{V}^{\ast}(v^{\lambda}_{i})=\sum_{i\neq j}a_{i}i_{V}^{\ast}(v^{\lambda}_{i}).

Therefore, iV(v1λ),,iV(vj1λ),iV(vj+1λ),,iV(vnλ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{j-1}),i_{V}^{\ast}(v^{\lambda}_{j+1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n}) are the direction vectors from the vertex iV(λ)i_{V}^{\ast}(\lambda). ∎

3.2. Case of iV(λ)i_{V}^{\ast}(\lambda) is not a vertex

Let λ\lambda be a vertex of the Delzant polytope Δ\Delta. In this section, we show that if the vertex λ\lambda of Δ\Delta does not satisfy the condition (1) in Definition 1.1, i.e. there is a nontrivial subspace WVλW^{\lambda}_{V} in the cone iV(𝒞λ)i_{V}^{\ast}(\mathcal{C}_{\lambda}), then Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular. From Proposition 3.5, this vertex λ\lambda satisfies the condition (2), i.e. 𝒥λ=\mathcal{J}_{\lambda}=\emptyset.

Let xx be a nonzero element in the nontrivial subspace WVλW^{\lambda}_{V}. Since xWVλx\in W^{\lambda}_{V}, we obtain (x)WVλ(-x)\in W^{\lambda}_{V}. Since the subspace WVλW^{\lambda}_{V} is in the cone iV(𝒞λ)i_{V}^{\ast}(\mathcal{C}_{\lambda}), there exist r1,,rn,s1,,sn0r_{1},\ldots,r_{n},s_{1},\ldots,s_{n}\geq 0 such that

x=j=1nrjiV(vjλ),(x)=j=1nsjiV(vjλ).x=\sum_{j=1}^{n}r_{j}i_{V}^{\ast}(v^{\lambda}_{j}),\;(-x)=\sum_{j=1}^{n}s_{j}i_{V}^{\ast}(v^{\lambda}_{j}).

Let Rj=rj+sjR_{j}=r_{j}+s_{j} for j=1,,nj=1,\ldots,n. Then, we obtain

(3.1) 0=x+(x)=j=1n(rj+sj)iV(vjλ)=j=1nRjiV(vjλ).0=x+(-x)=\sum_{j=1}^{n}(r_{j}+s_{j})i_{V}^{\ast}(v^{\lambda}_{j})=\sum_{j=1}^{n}R_{j}i_{V}^{\ast}(v^{\lambda}_{j}).

From Lemma 3.1, we may assume that the vectors iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) are linearly independent.

Lemma 3.7.

Assume that the vectors iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) are linearly independent. If iV(λ)i_{V}^{\ast}(\lambda) is not a vertex of the polytope iV(Δ)i_{V}^{\ast}(\Delta), then Rn0R_{n}\neq 0 in Equation 3.1.

Proof.

Assume on the contrary that Rn=0R_{n}=0. Then, from Equation 3.1 we obtain

0=j=1n1RjiV(vjλ).0=\sum_{j=1}^{n-1}R_{j}i_{V}^{\ast}(v^{\lambda}_{j}).

Since the vectors iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) are linearly independent, we obtain R1==Rn1=0R_{1}=\cdots=R_{n-1}=0. Since Rj=rj+sjR_{j}=r_{j}+s_{j} and rj,sj0r_{j},s_{j}\geq 0, we obtain rj=sj=0r_{j}=s_{j}=0 for any j=1,,nj=1,\ldots,n, i.e. x=0x=0. This is contradiction to the assumption that xx is a nonzero element in the subspace WVλW^{\lambda}_{V}. ∎

Lemma 3.8.

Assume that the vectors iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) are linearly independent. If iV(λ)i_{V}^{\ast}(\lambda) is not a vertex of the polytope iV(Δ)i_{V}^{\ast}(\Delta), then unλ,q0\langle u^{\lambda}_{n},q\rangle\neq 0.

Proof.

From Lemma 3.7, Rn0R_{n}\neq 0. Then, Equation 3.1 follows

iV(vnλ)=1Rnj=1n1RjiV(vjλ).i_{V}^{\ast}(v^{\lambda}_{n})=-\frac{1}{R_{n}}\sum_{j=1}^{n-1}R_{j}i_{V}^{\ast}(v^{\lambda}_{j}).

Since from Lemma 2.6 we calculate

j=1n1ujλ,qiV(vjλ)\displaystyle\sum_{j=1}^{n-1}\langle u^{\lambda}_{j},q\rangle i_{V}^{\ast}(v^{\lambda}_{j}) =unλ,qiV(vnλ)\displaystyle=-\langle u^{\lambda}_{n},q\rangle i_{V}^{\ast}(v^{\lambda}_{n})
=unλ,q1Rnj=1n1RjiV(vjλ)\displaystyle=\langle u^{\lambda}_{n},q\rangle\frac{1}{R_{n}}\sum_{j=1}^{n-1}R_{j}i_{V}^{\ast}(v^{\lambda}_{j})
=j=1n1unλ,qRjRniV(vjλ),\displaystyle=\sum_{j=1}^{n-1}\langle u^{\lambda}_{n},q\rangle\frac{R_{j}}{R_{n}}i_{V}^{\ast}(v^{\lambda}_{j}),

we obtain

j=1n1(ujλ,qunλ,qRjRn)iV(vjλ)=0.\sum_{j=1}^{n-1}\left(\langle u^{\lambda}_{j},q\rangle-\langle u^{\lambda}_{n},q\rangle\frac{R_{j}}{R_{n}}\right)i_{V}^{\ast}(v^{\lambda}_{j})=0.

Since we assume that the vectors iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) are linearly independent, we obtain

(3.2) u1λ,q=unλ,qR1Rn,,un1λ,q=unλ,qRn1Rn.\langle u^{\lambda}_{1},q\rangle=\langle u^{\lambda}_{n},q\rangle\frac{R_{1}}{R_{n}},\ldots,\langle u^{\lambda}_{n-1},q\rangle=\langle u^{\lambda}_{n},q\rangle\frac{R_{n-1}}{R_{n}}.

Since Rn0R_{n}\neq 0 means that Rn>0R_{n}>0, we say that

(3.3) R1Rn0,,Rn1Rn0.\frac{R_{1}}{R_{n}}\geq 0,\ldots,\frac{R_{n-1}}{R_{n}}\geq 0.

Assume on the contrary that unλ,q=0\langle u^{\lambda}_{n},q\rangle=0. Then, from Equation 3.2, we obtain u1λ,q==un1λ,q=0\langle u^{\lambda}_{1},q\rangle=\cdots=\langle u^{\lambda}_{n-1},q\rangle=0. Since u1λ,,unλu^{\lambda}_{1},\ldots,u^{\lambda}_{n} form a \mathbb{Z}-basis of n\mathbb{Z}^{n}, if u1λ,q==unλ,q=0\langle u^{\lambda}_{1},q\rangle=\cdots=\langle u^{\lambda}_{n},q\rangle=0, then q=0q=0. This is contradiction to the assumption that qq is a basis of the orthogonal subspace VV^{\perp} to the (n1)(n-1)-dimensional subspace VV in 𝔱nn\mathfrak{t}^{n}\cong\mathbb{R}^{n}. Therefore, unλ,q0\langle u^{\lambda}_{n},q\rangle\neq 0. ∎

Lemma 3.9.

If iV(λ)i_{V}^{\ast}(\lambda) is not a vertex of the polytope iV(Δ)i_{V}^{\ast}(\Delta), then λ+λ0=\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset or λλ0=\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset.

Proof.

From Lemma 3.1, we may assume that the vectors iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) are linearly independent.

By Lemma 3.8, we obtain unλ,q0\langle u^{\lambda}_{n},q\rangle\neq 0. Moreover, since Equation 3.2 and Equation 3.3, if nλ+λ0n\in\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}, then 1,,n1λ+1,\ldots,n-1\in\mathcal{I}^{+}_{\lambda}, i.e. λ0=\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}=\emptyset. Similarly, if nλλ0n\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}, then 1,,n1λ1,\ldots,n-1\in\mathcal{I}^{-}_{\lambda}, i.e. λ+0=\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}=\emptyset. ∎

From Lemma 3.9 and Lemma 2.8, we obtain the following:

Corollary 3.10.

If iV(λ)i_{V}^{\ast}(\lambda) is not a vertex of the polytope iV(Δ)i_{V}^{\ast}(\Delta), then Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular.

3.3. Case of 𝒥λ=\mathcal{J}_{\lambda}=\emptyset and iV(λ)i_{V}^{\ast}(\lambda) is a vertex

Let λ\lambda be a vertex of the Delzant polytope Δ\Delta. In this section, we consider the case when the vertex λ\lambda of Δ\Delta is a good vertex with respect to the map iV:(𝔱n)(𝔱n1)i_{V}^{\ast}:(\mathfrak{t}^{n})^{\ast}\to(\mathfrak{t}^{n-1})^{\ast} in the sense of Definition 1.1. In this case, we suppose the following:

Assumption 3.11.

The vectors iV(v1λ),,iV(vn1λ)ni_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1})\in\mathbb{Z}^{n} are linearly independent. Moreover, we have

iV({i=1naiviλai0})={j=1n1bjiV(vjλ)bj0}.i_{V}^{\ast}\left(\left\{\sum_{i=1}^{n}a_{i}v^{\lambda}_{i}\mid a_{i}\geq 0\right\}\right)=\left\{\sum_{j=1}^{n-1}b_{j}i_{V}^{\ast}(v^{\lambda}_{j})\mid b_{j}\geq 0\right\}.

Assumption 3.11 implies that iV(λ)i_{V}^{\ast}(\lambda) is a vertex of the polytope iV(Δ)i_{V}^{\ast}(\Delta). Note that if the polytope iV(Δ)i_{V}^{\ast}(\Delta) is simple, then Assumption 3.11 holds.

Lemma 3.12.

Under Assumption 3.11, if the vertex λ\lambda of Δ\Delta is a good vertex with respect to the map iV:(𝔱n)(𝔱n1)i_{V}^{\ast}:(\mathfrak{t}^{n})^{\ast}\to(\mathfrak{t}^{n-1})^{\ast}, then we obtain

λ+={1,,n1},λλ0={n}\mathcal{I}^{+}_{\lambda}=\{1,\ldots,n-1\},\;\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\{n\}

or

λ+λ0={n},λ={1,,n1}.\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\{n\},\;\mathcal{I}^{-}_{\lambda}=\{1,\ldots,n-1\}.
Proof.

From Assumption 3.11, there exist S1,,Sn10S_{1},\ldots,S_{n-1}\in\mathbb{R}_{\geq 0} such that

(3.4) iV(vnλ)=j=1n1SjiV(vjλ).i_{V}^{\ast}(v^{\lambda}_{n})=\sum_{j=1}^{n-1}S_{j}i_{V}^{\ast}(v^{\lambda}_{j}).

From Lemma 2.6, we calculate

0\displaystyle 0 =j=1nujλ,qiV(vjλ)\displaystyle=\sum_{j=1}^{n}\langle u^{\lambda}_{j},q\rangle i_{V}^{\ast}(v^{\lambda}_{j})
=j=1n1ujλ,qiV(vjλ)+unλ,qiV(vnλ)\displaystyle=\sum_{j=1}^{n-1}\langle u^{\lambda}_{j},q\rangle i_{V}^{\ast}(v^{\lambda}_{j})+\langle u^{\lambda}_{n},q\rangle i_{V}^{\ast}(v^{\lambda}_{n})
=j=1n1(ujλ,q+Sjunλ,q)iV(vjλ).\displaystyle=\sum_{j=1}^{n-1}\left(\langle u^{\lambda}_{j},q\rangle+S_{j}\langle u^{\lambda}_{n},q\rangle\right)i_{V}^{\ast}(v^{\lambda}_{j}).

Since iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) are linearly independent, we obtain

(3.5) u1λ,q=S1unλ,q,,un1λ,q=Sn1unλ,q.\langle u^{\lambda}_{1},q\rangle=-S_{1}\langle u^{\lambda}_{n},q\rangle,\ldots,\langle u^{\lambda}_{n-1},q\rangle=-S_{n-1}\langle u^{\lambda}_{n},q\rangle.

Note that if Sj=0S_{j}=0, then ujλ,q=0\langle u^{\lambda}_{j},q\rangle=0.

Assume on the contrary that unλ,q=0\langle u^{\lambda}_{n},q\rangle=0. If unλ,q=0\langle u^{\lambda}_{n},q\rangle=0, then we obtain

(3.6) u1λ,q==un1λ,q=0\langle u^{\lambda}_{1},q\rangle=\cdots=\langle u^{\lambda}_{n-1},q\rangle=0

from Equation 3.5. Since the vectors u1λ,,unλu^{\lambda}_{1},\ldots,u^{\lambda}_{n} form a \mathbb{Z}-basis of n\mathbb{Z}^{n}, Equation 3.6 means that q=0q=0, i.e. V={0}V^{\perp}=\{0\}. This is contradiction to the assumption that VV^{\perp} is the orthogonal subspace to the (n1)(n-1)-dimensional subspace VV in 𝔱nn\mathfrak{t}^{n}\cong\mathbb{R}^{n}. Therefore, unλ,q0\langle u^{\lambda}_{n},q\rangle\neq 0.

Since unλ,q0\langle u^{\lambda}_{n},q\rangle\neq 0, we obtain nλ0n\not\in\mathcal{I}^{0}_{\lambda}, i.e. nλ+λ0n\in\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda} or nλλ0n\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}. From Equation 3.5, if nλ+λ0n\in\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}, then 1,,n1λ1,\ldots,n-1\in\mathcal{I}^{-}_{\lambda}. Similarly, if nλλ0n\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}, then 1,,n1λ+1,\ldots,n-1\in\mathcal{I}^{+}_{\lambda}. ∎

By the definition of the pullback iV:(𝔱n)(𝔱n1)i_{V}^{\ast}:(\mathfrak{t}^{n})^{\ast}\to(\mathfrak{t}^{n-1})^{\ast}, it is clear that the vectors iV(v1λ),,iV(vnλ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n}) are integral vectors.

Lemma 3.13.

Under Assumption 3.11, if the vectors iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) form a \mathbb{Z}-basis of n1\mathbb{Z}^{n-1}, then unλ,q=±1\langle u^{\lambda}_{n},q\rangle=\pm 1.

Proof.

Since the vectors iV(v1λ),,iV(vn1λ)n1i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1})\in\mathbb{Z}^{n-1} form a \mathbb{Z}-basis of n1\mathbb{Z}^{n-1} and iV(vnλ)n1i_{V}^{\ast}(v^{\lambda}_{n})\in\mathbb{Z}^{n-1}, there exist Q1,,Qn1Q_{1},\ldots,Q_{n-1}\in\mathbb{Z} such that

iV(vnλ)=j=1n1QjiV(vjλ).i_{V}^{\ast}(v^{\lambda}_{n})=\sum_{j=1}^{n-1}Q_{j}i_{V}^{\ast}(v^{\lambda}_{j}).

Moreover, since iV(vnλ)iV(𝒞λ)i_{V}^{\ast}(v^{\lambda}_{n})\in i_{V}^{\ast}(\mathcal{C}_{\lambda}) from Assumption 3.11, iV(vnλ)i_{V}^{\ast}(v^{\lambda}_{n}) can be expressed as the form in Equation 3.4. From these equations, we obtain

j=1n1(QjSj)iV(vjλ)=0.\sum_{j=1}^{n-1}\left(Q_{j}-S_{j}\right)i_{V}^{\ast}(v^{\lambda}_{j})=0.

Since the vectors iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) are linearly independent,

(3.7) Q1=S1,,Qn1=Sn1.Q_{1}=S_{1},\ldots,Q_{n-1}=S_{n-1}.

Since S1,,Sn10S_{1},\ldots,S_{n-1}\geq 0, we have Q1,,Qn10Q_{1},\ldots,Q_{n-1}\geq 0. Thus, we obtain Q1,,Qn10Q_{1},\ldots,Q_{n-1}\in\mathbb{Z}_{\geq 0}.

From Equation 3.5 and Equation 3.7, we obtain

(3.8) u1λ,q=Q1unλ,q,,un1λ,q=Qn1unλ,q.\langle u^{\lambda}_{1},q\rangle=-Q_{1}\langle u^{\lambda}_{n},q\rangle,\ldots,\langle u^{\lambda}_{n-1},q\rangle=-Q_{n-1}\langle u^{\lambda}_{n},q\rangle.

From Equation 3.8, we obtain

[u1λunλ]tq=[u1λ,qunλ,q]=unλ,q[Q1Qn11]{{}^{t}\!\left[\begin{matrix}u^{\lambda}_{1}&\cdots&u^{\lambda}_{n}\end{matrix}\right]}q=\left[\begin{matrix}\langle u^{\lambda}_{1},q\rangle\\ \vdots\\ \langle u^{\lambda}_{n},q\rangle\end{matrix}\right]=\langle u^{\lambda}_{n},q\rangle\left[\begin{matrix}-Q_{1}\\ \vdots\\ -Q_{n-1}\\ 1\end{matrix}\right]

and since [u1λunλ]t[v1λvnλ]=En{{}^{t}[u^{\lambda}_{1}\;\cdots\;u^{\lambda}_{n}]}[v^{\lambda}_{1}\;\cdots\;v^{\lambda}_{n}]=E_{n}, we obtain

q=unλ,q[v1λvnλ][Q1Qn11].q=\langle u^{\lambda}_{n},q\rangle\left[\begin{matrix}v^{\lambda}_{1}&\cdots&v^{\lambda}_{n}\end{matrix}\right]\left[\begin{matrix}-Q_{1}\\ \vdots\\ -Q_{n-1}\\ 1\end{matrix}\right].

Since the vector qnq\in\mathbb{Z}^{n} is primitive, we obtain unλ,q=±1\langle u^{\lambda}_{n},q\rangle=\pm 1. ∎

From Lemma 2.9, Lemma 3.12 and Lemma 3.13, we obtain the following:

Corollary 3.14.

Assume that the vertex λ\lambda of the Delzant polytope Δ\Delta is a good vertex with respect to the map iV:(𝔱n)(𝔱n1)i_{V}^{\ast}:(\mathfrak{t}^{n})^{\ast}\to(\mathfrak{t}^{n-1})^{\ast}. Under Assumption 3.11, if there exist distinct j1,,jn1{1,,n}j_{1},\ldots,j_{n-1}\in\{1,\ldots,n\} such that the vectors iV(vj1λ),,iV(vjn1λ)n1i_{V}^{\ast}(v^{\lambda}_{j_{1}}),\ldots,i_{V}^{\ast}(v^{\lambda}_{j_{n-1}})\in\mathbb{Z}^{n-1} form a \mathbb{Z}-basis of n1\mathbb{Z}^{n-1}, then Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular.

3.4. The Proof of the first Main Theorem

In this section, we give a proof of the first main result.

Theorem 3.15.

If the Delzant polytope Δ\Delta is good with respect to the map iVi_{V}^{\ast}, then the rank of the Jacobian matrix DfλDf^{\lambda} is equal to one at any point for any vertex λ\lambda of Δ\Delta. In particular, C(V)¯\overline{C(V)} is a smooth complex hypersurface in the toric manifold XX corresponding to Δ\Delta.

Proof.

It is sufficient to consider each vertex of the Delznt polytope Δ\Delta.

If a vertex λ\lambda of Δ\Delta is not a good vertex with respect to the map iVi_{V}^{\ast}, then the vertex λ\lambda of Δ\Delta does not satisfy the condition (1) or (2) in Definition 1.1. If λ\lambda does not satisfy the condition (2), i.e. 𝒥λ\mathcal{J}_{\lambda}\neq\emptyset, then Corollary 3.4 tells us that Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular. If λ\lambda does not satisfy the condition (1), then Corollary 3.10 tells us that Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular. Thus, if a vertex λ\lambda of Δ\Delta is not a good vertex with respect to the map iVi_{V}^{\ast}, then Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular.

Since we assume that the Delzant polytope Δ\Delta is good with respect to the map iVi_{V}^{\ast}, if a vertex λ\lambda of Δ\Delta is a good vertex with respect to the map iVi_{V}^{\ast}, then Assumption 3.11 holds for such λ\lambda and there exist distinct j1,,jn1{1,,n}j_{1},\ldots,j_{n-1}\in\{1,\ldots,n\} such that the vectors iV(vj1λ),,iV(vjn1λ)n1i_{V}^{\ast}(v^{\lambda}_{j_{1}}),\ldots,i_{V}^{\ast}(v^{\lambda}_{j_{n-1}})\in\mathbb{Z}^{n-1} form a \mathbb{Z}-basis of n1\mathbb{Z}^{n-1}. In Corollary 3.14, we show that Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular in this case.

Therefore, if Δ\Delta is good with respect to the map iVi_{V}^{\ast}, then Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular for any vertex λ\lambda of Δ\Delta, i.e. C(V)¯\overline{C(V)} is a smooth complex hypersurface in the toric manifold XX. ∎

4. From Torus-equivariantly Embedded Toric Hypersurfaces to Delzant Polytopes

In this section, we show the second main result (Theorem 4.13).

Assume that the rank of the Jacobian matrix DfλDf^{\lambda} is equal to one at any point of the zero locus of fλf^{\lambda} for any vertex λ\lambda of the Delzant polytope, i.e. C(V)¯\overline{C(V)} is nonsingular. In this case, as we show in Proposition 2.5, C(V)¯\overline{C(V)} is toric with respect to the Hamiltonian Tn1T^{n-1}-action defined in [Yam24b] with the moment map μ¯=iVμC(V)¯\overline{\mu}=i_{V}^{\ast}\circ\mu\mid_{\overline{C(V)}}. Moreover, we have the following:

Theorem 4.1 ([Yam24b, Theorem 4.20]).

Assume that C(V)¯\overline{C(V)} is a complex submanifold in the toric manifold XX. Then, we obtain

μ¯(C(V)¯)=iV(Δ),\overline{\mu}(\overline{C(V)})=i_{V}^{\ast}(\Delta),

where Δ\Delta is the Delzant polytope of the toric manifold XX.

Recall that if C(V)¯\overline{C(V)} is a complex submanifold, then C(V)¯\overline{C(V)} is a symplectic toric manifold with respect to the Tn1T^{n-1}-action on C(V)¯\overline{C(V)} (Proposition 2.5). By the Delzant correspondence, the moment polytope of the submanifold C(V)¯\overline{C(V)} should be Delzant, i.e. we obtain the following:

Corollary 4.2.

Assume that C(V)¯\overline{C(V)} is a complex submanifold in the toric manifold XX. Then, the polytope iV(Δ)i_{V}^{\ast}(\Delta) is a Delzant polytope.

In particular, the polytope iV(Δ)i_{V}^{\ast}(\Delta) is simple, i.e. Assumption 3.11 holds under the assumption that C(V)¯\overline{C(V)} is a smooth complex hypersurface.

To show the second main result, we have to consider the case when the vertex λ\lambda of Δ\Delta is good with respect to the map iVi_{V}^{\ast}.

Hereafter, we assume that the vertex λ\lambda of Δ\Delta is good with respect to the map iVi_{V}^{\ast}. If C(V)¯\overline{C(V)} is a complex hypersurface, then from Lemma 3.12, we may further assume that λ+={1,,n1}\mathcal{I}^{+}_{\lambda}=\{1,\ldots,n-1\} and λλ0={n}\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\{n\} under the Assumption 3.11.

Lemma 4.3.

Assume that the vertex λ\lambda of Δ\Delta is good with respect to the map iVi_{V}^{\ast}. Under the Assumption 3.11, we obtain λ+λ0\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}\neq\emptyset and λλ0\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}\neq\emptyset.

Proof.

From Lemma 3.12, we may assume that λ+={1,,n1}\mathcal{I}^{+}_{\lambda}=\{1,\ldots,n-1\} and λλ0={n}\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\{n\}.

Assume on the contrary that λ+λ0=\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\emptyset, i.e. u1λ,q==un1λ,q=0\langle u^{\lambda}_{1},q\rangle=\cdots=\langle u^{\lambda}_{n-1},q\rangle=0. From Lemma 2.6, we obtain

0=j=1nujλ,qiV(vjλ)=unλ,qiV(vnλ).0=\sum_{j=1}^{n}\langle u^{\lambda}_{j},q\rangle i_{V}^{\ast}(v^{\lambda}_{j})=\langle u^{\lambda}_{n},q\rangle i_{V}^{\ast}(v^{\lambda}_{n}).

Since unλ,q0\langle u^{\lambda}_{n},q\rangle\neq 0, we obtain iV(vnλ)=0i_{V}^{\ast}(v^{\lambda}_{n})=0, which is contradiction to the assumption that the vertex λ\lambda of Δ\Delta is good with respect to the map iVi_{V}^{\ast}. ∎

Lemma 4.4.

Assume that the vertex λ\lambda of Δ\Delta is good with respect to the map iVi_{V}^{\ast} and λ+={1,,n1}\mathcal{I}^{+}_{\lambda}=\{1,\ldots,n-1\} and λλ0={n}\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\{n\}. If C(V)¯\overline{C(V)} is nonsingular, then we obtain at least one of the following:

  1. (1)

    there exists i0λ+i_{0}\in\mathcal{I}^{+}_{\lambda} such that ui0λ,q=1\langle u^{\lambda}_{i_{0}},q\rangle=1 and ujλ,q=0\langle u^{\lambda}_{j},q\rangle=0 for any jλ+{i0}j\in\mathcal{I}^{+}_{\lambda}\setminus\{i_{0}\},

  2. (2)

    unλ,q=1\langle u^{\lambda}_{n},q\rangle=-1.

Proof.

From the defining equation for Cλ(V)¯\overline{C_{\lambda}(V)}, we obtain that

fλ(zλ)=jλ+λ0(zjλ)ujλ,qjλλ0(zjλ)ujλ,q.f^{\lambda}(z^{\lambda})=\prod_{j\in\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{\langle u^{\lambda}_{j},q\rangle}-\prod_{j\in\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{-\langle u^{\lambda}_{j},q\rangle}.

Note that λ+λ0\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}\neq\emptyset and λλ0\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}\neq\emptyset from Lemma 4.3. Assume that λ+={1,,n1},λλ0={n}\mathcal{I}^{+}_{\lambda}=\{1,\ldots,n-1\},\;\mathcal{I}^{-}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\{n\}. From this equation, we obtain

fλziλ={uiλ,qjλ+λ0(zjλ)ujλ,qδijif i=1,,n1,unλ,q(znλ)unλ,q1if i=n.\frac{\partial f^{\lambda}}{\partial z^{\lambda}_{i}}=\begin{cases*}\langle u^{\lambda}_{i},q\rangle\prod_{j\in\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}}(z^{\lambda}_{j})^{\langle u^{\lambda}_{j},q\rangle-\delta_{ij}}&if $i=1,\ldots,n-1$,\\ \langle u^{\lambda}_{n},q\rangle(z^{\lambda}_{n})^{-\langle u^{\lambda}_{n},q\rangle-1}&if $i=n$.\end{cases*}

If C(V)¯\overline{C(V)} is nonsingular, then there exists i0{1,,n}i_{0}\in\{1,\ldots,n\} such that

fλzi0λ(o)0\frac{\partial f^{\lambda}}{\partial z^{\lambda}_{i_{0}}}(o)\neq 0

at the origin oφλ(Uλ)Cλ(V)¯o\in\varphi_{\lambda}(U_{\lambda})\cap\overline{C_{\lambda}(V)}. If such i0λ+i_{0}\in\mathcal{I}^{+}_{\lambda}, then we obtain ujλ,qδi0j=0\langle u^{\lambda}_{j},q\rangle-\delta_{i_{0}j}=0 for j=1,,n1j=1,\ldots,n-1, i.e. ui0λ,q=1\langle u^{\lambda}_{i_{0}},q\rangle=1 and ujλ,q=0\langle u^{\lambda}_{j},q\rangle=0 for any jλ+{i0}j\in\mathcal{I}^{+}_{\lambda}\setminus\{i_{0}\}. If such i0λ+i_{0}\notin\mathcal{I}^{+}_{\lambda}, i.e. i0=ni_{0}=n, then unλ,q=1\langle u^{\lambda}_{n},q\rangle=-1. ∎

Note that we obtain the similar result to Lemma 4.4 when we assume that λ+λ0={n}\mathcal{I}^{+}_{\lambda}\setminus\mathcal{I}^{0}_{\lambda}=\{n\} and λ={1,,n1}\mathcal{I}^{-}_{\lambda}=\{1,\ldots,n-1\}.

4.1. The first case in Lemma 4.4

We consider the first case in Lemma 4.4. Hereafter, we assume that u1λ,q=1\langle u^{\lambda}_{1},q\rangle=1 and u2λ,q==un1λ,q=0\langle u^{\lambda}_{2},q\rangle=\cdots=\langle u^{\lambda}_{n-1},q\rangle=0 for simplicity when we consider the first case in Lemma 4.4.

Lemma 4.5.

Assume that the vertex λ\lambda of Δ\Delta is good with respect to the map iVi_{V}^{\ast}. Moreover, assume that u1λ,q=1\langle u^{\lambda}_{1},q\rangle=1 and u2λ,q==un1λ,q=0\langle u^{\lambda}_{2},q\rangle=\cdots=\langle u^{\lambda}_{n-1},q\rangle=0. Then, we obtain

(4.1) iV(v1λ)=unλ,qiV(vnλ).i_{V}^{\ast}(v^{\lambda}_{1})=-\langle u^{\lambda}_{n},q\rangle i_{V}^{\ast}(v^{\lambda}_{n}).
Proof.

Since we assume that u1λ,q=1\langle u^{\lambda}_{1},q\rangle=1 and u2λ,q==un1λ,q=0\langle u^{\lambda}_{2},q\rangle=\cdots=\langle u^{\lambda}_{n-1},q\rangle=0, we obtain

j=1nujλ,qiV(vjλ)=iV(v1λ)+unλ,qiV(vnλ).\sum_{j=1}^{n}\langle u^{\lambda}_{j},q\rangle i_{V}^{\ast}(v^{\lambda}_{j})=i_{V}^{\ast}(v^{\lambda}_{1})+\langle u^{\lambda}_{n},q\rangle i_{V}^{\ast}(v^{\lambda}_{n}).

From Lemma 2.6, the left hand side of this equation is equal to zero. Therefore, we obtain Equation 4.1. ∎

Remark 4.6.

If λ\lambda is good, Assumption 3.11 holds, and Cλ(V)¯\overline{C_{\lambda}(V)} is nonsingular, then we have iV(v1λ)=unλ,qiV(vnλ)i_{V}^{\ast}(v^{\lambda}_{1})=-\langle u^{\lambda}_{n},q\rangle i_{V}^{\ast}(v^{\lambda}_{n}) with unλ,q>0-\langle u^{\lambda}_{n},q\rangle\in\mathbb{Z}_{>0}. Therefore, we may use the vector iV(vnλ)i_{V}^{\ast}(v^{\lambda}_{n}) instead of iV(v1λ)i_{V}^{\ast}(v^{\lambda}_{1}) in Assumption 3.11, i.e. we may assume the following:

  • the vectors iV(v2λ),,iV(vnλ)i_{V}^{\ast}(v^{\lambda}_{2}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n}) are linearly independent,

  • we have

    iV({i=1naiviλai0})={j=2nbjiV(vjλ)bj0}.i_{V}^{\ast}\left(\left\{\sum_{i=1}^{n}a_{i}v^{\lambda}_{i}\mid a_{i}\geq 0\right\}\right)=\left\{\sum_{j=2}^{n}b_{j}i_{V}^{\ast}(v^{\lambda}_{j})\mid b_{j}\geq 0\right\}.
Lemma 4.7.

Assume the same as in Lemma 4.5. Then, we obtain

det[p1pn1q]=±q,qdet[iV(v2λ)iV(vnλ)].\det[p_{1}\;\cdots\;p_{n-1}\;q]=\pm\langle q,q\rangle\det[i_{V}^{\ast}(v^{\lambda}_{2})\;\cdots\;i_{V}^{\ast}(v^{\lambda}_{n})].
Proof.

Since from Lemma 4.5, we obtain

[p1tpn1tqt][v1λvnλ]\displaystyle\left[\begin{matrix}{{}^{t}p_{1}}\\ \vdots\\ {{}^{t}p_{n-1}}\\ {{}^{t}q}\end{matrix}\right]\left[v^{\lambda}_{1}\;\cdots\;v^{\lambda}_{n}\right] =[p1,v1λp1,vnλpn1,v1λpn1,vnλq,v1λq,vnλ]\displaystyle=\left[\begin{matrix}\langle p_{1},v^{\lambda}_{1}\rangle&\cdots&\langle p_{1},v^{\lambda}_{n}\rangle\\ \vdots&\ddots&\vdots\\ \langle p_{n-1},v^{\lambda}_{1}\rangle&\cdots&\langle p_{n-1},v^{\lambda}_{n}\rangle\\ \langle q,v^{\lambda}_{1}\rangle&\cdots&\langle q,v^{\lambda}_{n}\rangle\end{matrix}\right]
=[iV(v1λ)iV(vnλ)q,v1λq,vnλ]\displaystyle=\left[\begin{matrix}i_{V}^{\ast}(v^{\lambda}_{1})&\cdots&i_{V}^{\ast}(v^{\lambda}_{n})\\ \langle q,v^{\lambda}_{1}\rangle&\cdots&\langle q,v^{\lambda}_{n}\rangle\end{matrix}\right]
=[unλ,qiV(vnλ)iV(v2λ)iV(vnλ)q,v1λq,v2λq,vnλ],\displaystyle=\left[\begin{matrix}-\langle u^{\lambda}_{n},q\rangle i_{V}^{\ast}(v^{\lambda}_{n})&i_{V}^{\ast}(v^{\lambda}_{2})&\cdots&i_{V}^{\ast}(v^{\lambda}_{n})\\ \langle q,v^{\lambda}_{1}\rangle&\langle q,v^{\lambda}_{2}\rangle&\cdots&\langle q,v^{\lambda}_{n}\rangle\end{matrix}\right],

we obtain by using the cofactor expansion that

det[p1pn1q]det[v1λvnλ]\displaystyle\det[p_{1}\;\cdots\;p_{n-1}\;q]\det[v^{\lambda}_{1}\;\cdots\;v^{\lambda}_{n}]
=det[p1pn1q]tdet[v1λvnλ]\displaystyle=\det{{}^{t}[p_{1}\;\cdots\;p_{n-1}\;q]}\det[v^{\lambda}_{1}\;\cdots\;v^{\lambda}_{n}]
=det[unλ,qiV(vnλ)iV(v2λ)iV(vnλ)q,v1λq,v2λq,vnλ]\displaystyle=\det\left[\begin{matrix}-\langle u^{\lambda}_{n},q\rangle i_{V}^{\ast}(v^{\lambda}_{n})&i_{V}^{\ast}(v^{\lambda}_{2})&\cdots&i_{V}^{\ast}(v^{\lambda}_{n})\\ \langle q,v^{\lambda}_{1}\rangle&\langle q,v^{\lambda}_{2}\rangle&\cdots&\langle q,v^{\lambda}_{n}\rangle\end{matrix}\right]
=det[0iV(v2λ)iV(vnλ)q,v1λ+q,vnλunλ,qq,v2λq,vnλ]\displaystyle=\det\left[\begin{matrix}0&i_{V}^{\ast}(v^{\lambda}_{2})&\cdots&i_{V}^{\ast}(v^{\lambda}_{n})\\ \langle q,v^{\lambda}_{1}\rangle+\langle q,v^{\lambda}_{n}\rangle\langle u^{\lambda}_{n},q\rangle&\langle q,v^{\lambda}_{2}\rangle&\cdots&\langle q,v^{\lambda}_{n}\rangle\end{matrix}\right]
=±(q,v1λ+q,vnλunλ,q)det[iV(v2λ)iV(vnλ)].\displaystyle=\pm\left(\langle q,v^{\lambda}_{1}\rangle+\langle q,v^{\lambda}_{n}\rangle\langle u^{\lambda}_{n},q\rangle\right)\det[i_{V}^{\ast}(v^{\lambda}_{2})\;\cdots\;i_{V}^{\ast}(v^{\lambda}_{n})].

Since we assume u1λ,q=1\langle u^{\lambda}_{1},q\rangle=1 and u2λ,q==un1λ,q=0\langle u^{\lambda}_{2},q\rangle=\cdots=\langle u^{\lambda}_{n-1},q\rangle=0, we obtain

q,v1λ+q,vnλunλ,q=j=1nq,vjλujλ,q=q,q.\langle q,v^{\lambda}_{1}\rangle+\langle q,v^{\lambda}_{n}\rangle\langle u^{\lambda}_{n},q\rangle=\sum_{j=1}^{n}\langle q,v^{\lambda}_{j}\rangle\langle u^{\lambda}_{j},q\rangle=\langle q,q\rangle.

Thus, we obtain

±(q,v1λ+q,vnλunλ,q)det[iV(v2λ)iV(vnλ)]\displaystyle\pm\left(\langle q,v^{\lambda}_{1}\rangle+\langle q,v^{\lambda}_{n}\rangle\langle u^{\lambda}_{n},q\rangle\right)\det[i_{V}^{\ast}(v^{\lambda}_{2})\;\cdots\;i_{V}^{\ast}(v^{\lambda}_{n})]
=±q,qdet[iV(v2λ)iV(vnλ)].\displaystyle=\pm\langle q,q\rangle\det[i_{V}^{\ast}(v^{\lambda}_{2})\;\cdots\;i_{V}^{\ast}(v^{\lambda}_{n})].

Since v1λ,,vnλv^{\lambda}_{1},\ldots,v^{\lambda}_{n} form a basis of n\mathbb{Z}^{n}, det[v1λvnλ]=±1\det[v^{\lambda}_{1}\;\cdots\;v^{\lambda}_{n}]=\pm 1, and

det[p1pn1q]=±q,qdet[iV(v2λ)iV(vnλ)].\det[p_{1}\;\cdots\;p_{n-1}\;q]=\pm\langle q,q\rangle\det[i_{V}^{\ast}(v^{\lambda}_{2})\;\cdots\;i_{V}^{\ast}(v^{\lambda}_{n})].

4.2. The second case in Lemma 4.4

Next, we consider the second case in Lemma 4.4, i.e. we assume that unλ,q=1\langle u^{\lambda}_{n},q\rangle=-1 and λ+={1,,n1}\mathcal{I}^{+}_{\lambda}=\{1,\ldots,n-1\}.

Lemma 4.8.

Assume that the vertex λ\lambda of Δ\Delta is good with respect to the map iVi_{V}^{\ast} and that unλ,q=1\langle u^{\lambda}_{n},q\rangle=-1 and λ+={1,,n1}\mathcal{I}^{+}_{\lambda}=\{1,\ldots,n-1\}. Then, we obtain

(4.2) iV(vnλ)=j=1n1ujλ,qiV(vjλ).i_{V}^{\ast}(v^{\lambda}_{n})=\sum_{j=1}^{n-1}\langle u^{\lambda}_{j},q\rangle i_{V}^{\ast}(v^{\lambda}_{j}).
Proof.

From Lemma 2.6, we obtain

0=j=1nujλ,qiV(vjλ)=j=1n1ujλ,qiV(vjλ)iV(vnλ).\displaystyle 0=\sum_{j=1}^{n}\langle u^{\lambda}_{j},q\rangle i_{V}^{\ast}(v^{\lambda}_{j})=\sum_{j=1}^{n-1}\langle u^{\lambda}_{j},q\rangle i_{V}^{\ast}(v^{\lambda}_{j})-i_{V}^{\ast}(v^{\lambda}_{n}).

Thus, we obtain Equation 4.2. ∎

Lemma 4.9.

Assume that unλ,q=1\langle u^{\lambda}_{n},q\rangle=-1 and λ+={1,,n1}\mathcal{I}^{+}_{\lambda}=\{1,\ldots,n-1\}. Then, we obtain

det[p1pn1q]=±q,qdet[iV(v1λ)iV(vn1λ)].\det[p_{1}\;\cdots\;p_{n-1}\;q]=\pm\langle q,q\rangle\det[i_{V}^{\ast}(v^{\lambda}_{1})\;\cdots\;i_{V}^{\ast}(v^{\lambda}_{n-1})].
Proof.

Since we see

[p1tpn1tqt][v1λvnλ]=[iV(v1λ)iV(vnλ)q,v1λq,vnλ],\displaystyle\left[\begin{matrix}{{}^{t}p_{1}}\\ \vdots\\ {{}^{t}p_{n-1}}\\ {{}^{t}q}\end{matrix}\right]\left[v^{\lambda}_{1}\;\cdots\;v^{\lambda}_{n}\right]=\left[\begin{matrix}i_{V}^{\ast}(v^{\lambda}_{1})&\cdots&i_{V}^{\ast}(v^{\lambda}_{n})\\ \langle q,v^{\lambda}_{1}\rangle&\cdots&\langle q,v^{\lambda}_{n}\rangle\end{matrix}\right],

we obtain by using the cofactor expansion that

det[p1pn1q]det[v1λvnλ]\displaystyle\det[p_{1}\;\cdots\;p_{n-1}\;q]\det[v^{\lambda}_{1}\;\cdots\;v^{\lambda}_{n}]
=det[iV(v1λ)iV(vnλ)q,v1λq,vnλ]\displaystyle=\det\left[\begin{matrix}i_{V}^{\ast}(v^{\lambda}_{1})&\cdots&i_{V}^{\ast}(v^{\lambda}_{n})\\ \langle q,v^{\lambda}_{1}\rangle&\cdots&\langle q,v^{\lambda}_{n}\rangle\end{matrix}\right]
=det[iV(v1λ)iV(vn1λ)j=1n1ujλ,qiV(vjλ)q,v1λq,vn1λq,vnλ]\displaystyle=\det\left[\begin{matrix}i_{V}^{\ast}(v^{\lambda}_{1})&\cdots&i_{V}^{\ast}(v^{\lambda}_{n-1})&\sum_{j=1}^{n-1}\langle u^{\lambda}_{j},q\rangle i_{V}^{\ast}(v^{\lambda}_{j})\\ \langle q,v^{\lambda}_{1}\rangle&\cdots&\langle q,v^{\lambda}_{n-1}\rangle&\langle q,v^{\lambda}_{n}\rangle\end{matrix}\right]
=det[iV(v1λ)iV(vn1λ)0q,v1λq,vn1λq,vnλj=1n1q,vjλujλ,q]\displaystyle=\det\left[\begin{matrix}i_{V}^{\ast}(v^{\lambda}_{1})&\cdots&i_{V}^{\ast}(v^{\lambda}_{n-1})&0\\ \langle q,v^{\lambda}_{1}\rangle&\cdots&\langle q,v^{\lambda}_{n-1}\rangle&\langle q,v^{\lambda}_{n}\rangle-\sum_{j=1}^{n-1}\langle q,v^{\lambda}_{j}\rangle\langle u^{\lambda}_{j},q\rangle\end{matrix}\right]
=(q,vnλj=1n1q,vjλujλ,q)det[iV(v1λ)iV(vn1λ)].\displaystyle=\left(\langle q,v^{\lambda}_{n}\rangle-\sum_{j=1}^{n-1}\langle q,v^{\lambda}_{j}\rangle\langle u^{\lambda}_{j},q\rangle\right)\det[i_{V}^{\ast}(v^{\lambda}_{1})\;\cdots\;i_{V}^{\ast}(v^{\lambda}_{n-1})].

Since we assume unλ,q=1\langle u^{\lambda}_{n},q\rangle=-1, we obtain

q,vnλj=1n1q,vjλujλ,q=j=1nq,vjλujλ,q=q,q.\langle q,v^{\lambda}_{n}\rangle-\sum_{j=1}^{n-1}\langle q,v^{\lambda}_{j}\rangle\langle u^{\lambda}_{j},q\rangle=-\sum_{j=1}^{n}\langle q,v^{\lambda}_{j}\rangle\langle u^{\lambda}_{j},q\rangle=-\langle q,q\rangle.

Thus, we obtain

(q,vnλj=1n1q,vjλujλ,q)det[iV(v1λ)iV(vn1λ)]\displaystyle\left(\langle q,v^{\lambda}_{n}\rangle-\sum_{j=1}^{n-1}\langle q,v^{\lambda}_{j}\rangle\langle u^{\lambda}_{j},q\rangle\right)\det[i_{V}^{\ast}(v^{\lambda}_{1})\;\cdots\;i_{V}^{\ast}(v^{\lambda}_{n-1})]
=q,qdet[iV(v1λ)iV(vn1λ)].\displaystyle=-\langle q,q\rangle\det[i_{V}^{\ast}(v^{\lambda}_{1})\;\cdots\;i_{V}^{\ast}(v^{\lambda}_{n-1})].

Since v1λ,,vnλv^{\lambda}_{1},\ldots,v^{\lambda}_{n} form a basis of n\mathbb{Z}^{n}, det[v1λvnλ]=±1\det[v^{\lambda}_{1}\;\cdots\;v^{\lambda}_{n}]=\pm 1, and

det[p1pn1q]=±q,qdet[iV(v1λ)iV(vn1λ)].\det[p_{1}\;\cdots\;p_{n-1}\;q]=\pm\langle q,q\rangle\det[i_{V}^{\ast}(v^{\lambda}_{1})\;\cdots\;i_{V}^{\ast}(v^{\lambda}_{n-1})].

4.3. The Prood of the second Main Theorem

By calculating the determinant of the Gram matrix of [p1pn1q][p_{1}\;\cdots\;p_{n-1}\;q], we obtain the following:

Lemma 4.10.

For V=p1++pn1V=\mathbb{R}p_{1}+\cdots+\mathbb{R}p_{n-1} and V=qV^{\perp}=\mathbb{R}q, we obtain

det[p1pn1q]=±(q,qdet[p1,p1p1,pn1pn1,p1pn1,pn1])12.\det[p_{1}\;\cdots\;p_{n-1}\;q]=\pm\left(\langle q,q\rangle\det\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle\\ \vdots&\ddots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle\end{matrix}\right]\right)^{\frac{1}{2}}.
Proof.

Since the Gram matrix of [p1pn1q][p_{1}\;\cdots\;p_{n-1}\;q] is

[p1tpn1tqt][p1pn1q]\displaystyle\left[\begin{matrix}{{}^{t}p_{1}}\\ \vdots\\ {{}^{t}p_{n-1}}\\ {{}^{t}q}\end{matrix}\right][p_{1}\;\cdots\;p_{n-1}\;q] =[p1,p1p1,pn1p1,qpn1,p1pn1,pn1p1,qq,p1q,pn1q,q]\displaystyle=\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle&\langle p_{1},q\rangle\\ \vdots&\ddots&\vdots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle&\langle p_{1},q\rangle\\ \langle q,p_{1}\rangle&\cdots&\langle q,p_{n-1}\rangle&\langle q,q\rangle\end{matrix}\right]
=[p1,p1p1,pn10pn1,p1pn1,pn1000q,q],\displaystyle=\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle&0\\ \vdots&\ddots&\vdots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle&0\\ 0&\cdots&0&\langle q,q\rangle\end{matrix}\right],

we obtain

(det[p1pn1q])2\displaystyle(\det[p_{1}\;\cdots\;p_{n-1}\;q])^{2} =det([p1tpn1tqt][p1pn1q])\displaystyle=\det\left(\left[\begin{matrix}{{}^{t}p_{1}}\\ \vdots\\ {{}^{t}p_{n-1}}\\ {{}^{t}q}\end{matrix}\right][p_{1}\;\cdots\;p_{n-1}\;q]\right)
=det[p1,p1p1,pn10pn1,p1pn1,pn1000q,q]\displaystyle=\det\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle&0\\ \vdots&\ddots&\vdots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle&0\\ 0&\cdots&0&\langle q,q\rangle\end{matrix}\right]
=q,qdet[p1,p1p1,pn1pn1,p1pn1,pn1].\displaystyle=\langle q,q\rangle\det\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle\\ \vdots&\ddots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle\end{matrix}\right].

Since Gram matrices are positive definite, i.e. their determinant are nonnegative, we obtain the desired result. ∎

In our case, we obtain more about the determinant of Gram matrices. Though the following proposition is found in books [Mar03, Proposition 1.9.8] and [Ebe13, Proposition 1.2], we give a proof in order to make this paper self-contained.

Proposition 4.11.

Assume that p1,,pn1np_{1},\ldots,p_{n-1}\in\mathbb{Z}^{n} form a \mathbb{Z}-basis of Vnn1V\cap\mathbb{Z}^{n}\cong\mathbb{Z}^{n-1}. Let qnq\in\mathbb{Z}^{n} be a primitive basis of the orthogonal subspace to VV in n\mathbb{R}^{n}. For V=p1++pn1V=\mathbb{R}p_{1}+\cdots+\mathbb{R}p_{n-1} and V=qV^{\perp}=\mathbb{R}q, we obtain

det[p1,p1p1,pn1pn1,p1pn1,pn1]=q,q.\det\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle\\ \vdots&\ddots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle\end{matrix}\right]=\langle q,q\rangle.
Proof.

By the assumption, there exists a vector pnnp_{n}\in\mathbb{Z}^{n} such that p1,,pn1,pnp_{1},\ldots,p_{n-1},p_{n} form a \mathbb{Z}-basis of n\mathbb{Z}^{n}. Since we have

[p1tpn1tpnt][p1pn1q]\displaystyle\left[\begin{matrix}{{}^{t}p_{1}}\\ \vdots\\ {{}^{t}p_{n-1}}\\ {{}^{t}p_{n}}\end{matrix}\right]\left[p_{1}\;\cdots\;p_{n-1}\;q\right] =[p1,p1p1,pn1p1,qpn1,p1pn1,pn1pn1,qpn,p1pn,pn1pn,q]\displaystyle=\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle&\langle p_{1},q\rangle\\ \vdots&\ddots&\vdots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle&\langle p_{n-1},q\rangle\\ \langle p_{n},p_{1}\rangle&\cdots&\langle p_{n},p_{n-1}\rangle&\langle p_{n},q\rangle\end{matrix}\right]
=[p1,p1p1,pn10pn1,p1pn1,pn10pn,p1pn,pn1pn,q],\displaystyle=\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle&0\\ \vdots&\ddots&\vdots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle&0\\ \langle p_{n},p_{1}\rangle&\cdots&\langle p_{n},p_{n-1}\rangle&\langle p_{n},q\rangle\end{matrix}\right],

we obtain by the cofactor expansion that

det[p1pn1pn]det[p1pn1q]\displaystyle\det[p_{1}\;\cdots\;p_{n-1}\;p_{n}]\det[p_{1}\;\cdots\;p_{n-1}\;q]
=det[p1,p1p1,pn10pn1,p1pn1,pn10pn,p1pn,pn1pn,q]\displaystyle=\det\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle&0\\ \vdots&\ddots&\vdots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle&0\\ \langle p_{n},p_{1}\rangle&\cdots&\langle p_{n},p_{n-1}\rangle&\langle p_{n},q\rangle\end{matrix}\right]
=pn,qdet[p1,p1p1,pn1pn1,p1pn1,pn1].\displaystyle=\langle p_{n},q\rangle\det\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle\\ \vdots&\ddots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle\end{matrix}\right].

Since p1,,pnp_{1},\ldots,p_{n} form a \mathbb{Z}-basis of n\mathbb{Z}^{n}, we obtain det[p1pn1pn]=±1\det[p_{1}\;\cdots\;p_{n-1}\;p_{n}]=\pm 1. From Lemma 4.10, we obtain

(4.3) pn,q2det[p1,p1p1,pn1pn1,p1pn1,pn1]=q,q.\langle p_{n},q\rangle^{2}\det\left[\begin{matrix}\langle p_{1},p_{1}\rangle&\cdots&\langle p_{1},p_{n-1}\rangle\\ \vdots&\ddots&\vdots\\ \langle p_{n-1},p_{1}\rangle&\cdots&\langle p_{n-1},p_{n-1}\rangle\end{matrix}\right]=\langle q,q\rangle.

Since we have

[p1,qpn1,qpn,q]=[p1tpn1tpnt]q,\left[\begin{matrix}\langle p_{1},q\rangle\\ \vdots\\ \langle p_{n-1},q\rangle\\ \langle p_{n},q\rangle\end{matrix}\right]=\left[\begin{matrix}{{}^{t}p_{1}}\\ \vdots\\ {{}^{t}p_{n-1}}\\ {{}^{t}p_{n}}\end{matrix}\right]q,

and p1,,pnp_{1},\ldots,p_{n} form a \mathbb{Z}-basis of n\mathbb{Z}^{n}, we obtain

q\displaystyle q =[p1tpn1tpnt]1[p1,qpn1,qpn,q]\displaystyle=\left[\begin{matrix}{{}^{t}p_{1}}\\ \vdots\\ {{}^{t}p_{n-1}}\\ {{}^{t}p_{n}}\end{matrix}\right]^{-1}\left[\begin{matrix}\langle p_{1},q\rangle\\ \vdots\\ \langle p_{n-1},q\rangle\\ \langle p_{n},q\rangle\end{matrix}\right]
=[p1tpn1tpnt]1[00pn,q]\displaystyle=\left[\begin{matrix}{{}^{t}p_{1}}\\ \vdots\\ {{}^{t}p_{n-1}}\\ {{}^{t}p_{n}}\end{matrix}\right]^{-1}\left[\begin{matrix}0\\ \vdots\\ 0\\ \langle p_{n},q\rangle\end{matrix}\right]
=pn,q[p1tpn1tpnt]1[001].\displaystyle=\langle p_{n},q\rangle\left[\begin{matrix}{{}^{t}p_{1}}\\ \vdots\\ {{}^{t}p_{n-1}}\\ {{}^{t}p_{n}}\end{matrix}\right]^{-1}\left[\begin{matrix}0\\ \vdots\\ 0\\ 1\end{matrix}\right].

Since we assume that the vector qq is primitive, we obtain pn,q=±1\langle p_{n},q\rangle=\pm 1. Therefore, from Equation 4.3 we obtain the desired equation. ∎

From Lemma 4.10 and Proposition 4.11, we obtain the following:

Corollary 4.12.

Assume that p1,,pn1np_{1},\ldots,p_{n-1}\in\mathbb{Z}^{n} form a \mathbb{Z}-basis of Vnn1V\cap\mathbb{Z}^{n}\cong\mathbb{Z}^{n-1}. Assume that the vector qnq\in\mathbb{Z}^{n} is a primitive basis of the orthogonal subspace to VV in n\mathbb{R}^{n}. For V=p1++pn1V=\mathbb{R}p_{1}+\cdots+\mathbb{R}p_{n-1} and V=qV^{\perp}=\mathbb{R}q, we obtain

det[p1pn1q]=±q,q.\det[p_{1}\;\cdots\;p_{n-1}\;q]=\pm\langle q,q\rangle.

We use this fact to show that the second main result:

Theorem 4.13.

If the rank of the Jacobian matrix DfλDf^{\lambda} is equal to one at any point of the zero locus of fλf^{\lambda} for any vertex λ\lambda of the Delzant polytope Δ\Delta, i.e. C(V)¯\overline{C(V)} is nonsingular, then Δ\Delta is good with respect to the map iVi_{V}^{\ast}.

Proof.

It is sufficient to consider each vertex λ\lambda of the Delznt polytope Δ\Delta.

From Lemma 3.12 and Lemma 4.4, if the vertex λ\lambda of Δ\Delta is good with respect to the map iVi_{V}^{\ast}, then we obtain the two cases:

  1. (1)

    there exists i0λ+i_{0}\in\mathcal{I}^{+}_{\lambda} such that ui0λ,q=1\langle u^{\lambda}_{i_{0}},q\rangle=1 and ujλ,q=0\langle u^{\lambda}_{j},q\rangle=0 for any jλ+{i0}j\in\mathcal{I}^{+}_{\lambda}\setminus\{i_{0}\},

  2. (2)

    unλ,q=1\langle u^{\lambda}_{n},q\rangle=-1.

We first consider the first case. In this case, we can assume further that u1λ,q=1\langle u^{\lambda}_{1},q\rangle=1 and u2λ,q==un1λ,q=0\langle u^{\lambda}_{2},q\rangle=\cdots=\langle u^{\lambda}_{n-1},q\rangle=0. Under this assumption, Remark 4.6 shows that the vectors iV(v2λ),,iV(vnλ)i_{V}^{\ast}(v^{\lambda}_{2}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n}) are linearly independent and

iV({i=1naiviλai0})={j=2nbjiV(vjλ)bj0}.i_{V}^{\ast}\left(\left\{\sum_{i=1}^{n}a_{i}v^{\lambda}_{i}\mid a_{i}\geq 0\right\}\right)=\left\{\sum_{j=2}^{n}b_{j}i_{V}^{\ast}(v^{\lambda}_{j})\mid b_{j}\geq 0\right\}.

From Lemma 4.7 and Corollary 4.12, we obtain

det[iV(v2λ)iV(vnλ)]=±1,\det[i_{V}^{\ast}(v^{\lambda}_{2})\;\cdots i_{V}^{\ast}(v^{\lambda}_{n})]=\pm 1,

which shows that the vectors iV(v2λ),,iV(vnλ)i_{V}^{\ast}(v^{\lambda}_{2}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n}) form a \mathbb{Z}-basis of n1\mathbb{Z}^{n-1}.

We next consider the second case. In this case, from Lemma 4.9 and Corollary 4.12, we obtain

det[iV(v1λ)iV(vn1λ)]=±1,\det[i_{V}^{\ast}(v^{\lambda}_{1})\;\cdots i_{V}^{\ast}(v^{\lambda}_{n-1})]=\pm 1,

which shows that the vectors iV(v1λ),,iV(vn1λ)i_{V}^{\ast}(v^{\lambda}_{1}),\ldots,i_{V}^{\ast}(v^{\lambda}_{n-1}) form a \mathbb{Z}-basis of n1\mathbb{Z}^{n-1}.

Therefore, Δ\Delta is good with respect to the map iVi_{V}^{\ast}. ∎

References

  • [Ati82] M. F. Atiyah. Convexity and commuting Hamiltonians. Bull. London Math. Soc., 14(1):1–15, 1982.
  • [CDG03] David M. J. Calderbank, Liana David, and Paul Gauduchon. The Guillemin formula and Kähler metrics on toric symplectic manifolds. J. Symplectic Geom., 1(4):767–784, 2003.
  • [Del88] Thomas Delzant. Hamiltoniens périodiques et images convexes de l’application moment. Bull. Soc. Math. France, 116(3):315–339, 1988.
  • [Ebe13] Wolfgang Ebeling. Lattices and codes. Advanced Lectures in Mathematics. Springer Spektrum, Wiesbaden, third edition, 2013. A course partially based on lectures by Friedrich Hirzebruch.
  • [GS82] V. Guillemin and S. Sternberg. Convexity properties of the moment mapping. Invent. Math., 67(3):491–513, 1982.
  • [Gui94a] Victor Guillemin. Kaehler structures on toric varieties. J. Differential Geom., 40(2):285–309, 1994.
  • [Gui94b] Victor Guillemin. Moment maps and combinatorial invariants of Hamiltonian TnT^{n}-spaces, volume 122 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1994.
  • [Mar03] Jacques Martinet. Perfect lattices in Euclidean spaces, volume 327 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 2003.
  • [Yam24a] Kentaro Yamaguchi. Submanifolds with corners in Delzant polytopes associated to affine subspaces, 2024. preprint arXiv:2402.16884.
  • [Yam24b] Kentaro Yamaguchi. Torus-equivariantly embedded toric manifolds associated to affine subspaces, 2024. preprint arXiv:2306.15312, to appear in Osaka Journal of Mathematics.