Delzant type theorem for torus-equivariantly embedded toric hypersurfaces
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 embedding2020 Mathematics Subject Classification:
Primary 53C40; Secondary 53D20, 14M251. Introduction
Symplectic toric manifolds are -dimensional symplectic manifolds equipped with the effective Hamiltonian action of an -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 in the corresponding symplectic toric manifold can be identified with the complex torus .
In [Yam24b], we construct the closure of the -dimensional complex subtorus in the toric divior complement from the data of a -dimensional affine subspace . The closure can be expressed by the zero locus of polynomials for each vertex of the Delzant polytope, which are determined by the data of the Delzant polytope and the affine subspace. Note that the closure might have singularity at the intersection with the toric diviors . We show in [Yam24b, Theorem 4.20] that if the closure is a smooth -dimensional complex submanifold in the symplectic toric manifold , then the moment polytope for the Hamiltonian -dimensional torus action on coincides with the moment polytope for the Hamiltonian -dimensional subtorus action on . Moreover, the submanifold is a symplectic toric manifold with respect to the -dimensional torus action on . In particular, if is a complex submanifold, then the moment polytope for the Hamiltonian -dimensional subtorus action on is a Delzant polytope. We call the complex submanifold the torus-equivariantly embedded toric manifold.
Even though the moment polytope for the Hamiltonian -dimensional subtorus action on is a Delzant polytope as a convex polytope for some affine subspace , might not be a smooth complex submanifold. For example, let . For any -dimensional affine subspace in , the moment polytope for the Hamiltonian one-dimensional subtorus action associated with becomes an interval in , which can be seen as the Delzant polytope of . However, there are many affine subspaces such that the closure 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. .
Let be an -dimensional affine subspace in with rational slope, i.e. , where is the linear part of . Then, we may take a primitive -basis of and such that . We define the injective homomorphism by
where are the standard basis of . We introduce some notions concerning the pair of a Delzant polytope and an affine subspace.
Definition 1.1.
Let be a Delzant polytope and an -dimensional affine subspace in with rational slope. A vertex of the polytope is a good vertex with respect to the pullback if the vertex satisfies the two conditions:
-
(1)
is a vertex of the convex polytope ,
-
(2)
for the direction vectors of the edges from the vertex of the polytope , the vectors in are all nonzero.
Definition 1.2.
Let be an -dimensional affine subspace in with rational slope. A Delzant polytope is a good polytope with respect to the map if any good vertex with respect to the map satisfies the two conditions:
-
•
we can choose direction vectors such that are linearly independent and ,
-
•
the direction vectors from the vertex of form a -basis of .
Note that the first condition in Definition 1.2 are equivalent to the condition that the polytope is simple.
The closure can be expressed by the zero locus of a polynomial for each vertex of . Our first main result is as follows:
Theorem 1.3 (Theorem 3.15).
Let be a Delzant polytope and an -dimensional affine subspace in with rational slope. If the Delzant polytope is good with respect to the map , then the rank of the Jacobian matrix is equal to one at any point of the zero locus of for any vertex of the Delzant polytope . In particular, is a smooth complex hypersurface in the toric manifold .
Since complex hypersurfaces are symplectic toric manifolds with respect to the Hamiltonian -action, we obtain the following:
Corollary 1.4.
If the Delzant polytope of a toric manifold is good with respect to the map , then the convex polytope is a Delzant polytope of the complex hypersurface in .
We also show the converse of Theorem 1.3.
Theorem 1.5 (Theorem 4.13).
Let be an -dimensional affine subspace in with rational slope. If the rank of the Jacobian matrix is equal to one at any point of the zero locus of for any vertex of the Delzant polytope of the toric manifold , then the Delzant polytope is good with respect to the map .
Corollary 1.6.
Let be a Delzant polytope of a symplectic toric manifold . Then, is a torus-equivariantly embedded toric hypersurface in if and only if is a good Delzant polytope with respect to the map .
Corollary 1.6 characterizes the condition when 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 of the Delzant polytope does not satisfy the condition (2) in Definition 1.1, then the rank of the Jacobian matrix is equal to one. Next, in Section 3.2 we show that if the vertex satisfies the condition (2) but does not satisfy the condition (1) in Definition 1.1, then the rank of the Jacobian matrix is equal to one. Finally, in Section 3.3 we consider the case when the vertex is good with respect to the map . The detailed proof of Theorem 1.3 is given in Section 3.4.
1.3. Outline
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 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 in is Delzant if satisfies the following three properties:
-
•
simple; each vertex has edges,
-
•
rational; the direction vectors of the edges from any vertex can be chosen as integral vectors,
-
•
smooth; the vectors chosen as above form a -basis of .
Here, denotes the set of the vertices of .
Some literatures (for example, [CDG03]) define Delzant polytopes in terms of the inward pointing normal vectors to the 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:
holds for any vertex of (see [Yam24b, Lemma 3.10] for example).
Due to the Delzant construction, we obtain the corresponding symplectic toric manifold from the data of a given Delzant polytope . Here we give the expression of a system of the inhomogeneous coordinate charts on the corresponding symplectic toric manifold from the data of the Delzant polytope . We define two matrices and by
for any vertices .
Lemma 2.2.
From the data of a given Delzant polytope , we can construct an open covering of the corresponding toric manifold and a set of maps such that
for any vertices such that .
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 -dimensional symplectic toric manifold can be identified with a complex -dimensional torus.
2.2. Torus-equivariantly Embedded Toric Hypersurfaces
Let be primitive vectors which are linearly independent, and . We consider the -dimensional affine subspace in . Write as the linear part of . Assume that has a rational slope, i.e. we may assume that form a -basis of . Let be a primitive basis of the orthogonal subspace to in .
Through the log-affine coordinate , we define the complex subtorus (see [Yam24b, Proposition 4.2] for detail). Recall that the complement of toric diviors in a toric manifold can be identified with a complex -dimensional torus . The closure of in the toric manifold is expressed as follows:
Definition 2.3.
The closure of in the toric manifold is defined as , where is a zero locus of . Here, are defined by
where
Note that if , then . Similarly, if , then .
Remark 2.4.
The rank of the Jacobian matrix of might be equal to zero at some points in the toric diviors. In particular, might have a singularity at some points in the toric diviors.
Proposition 2.5.
If is a complex submanifold in the toric manifold , then is toric with respect to the -action on .
Proof.
We show that the -action on is effective, Hamiltonian. Define the injective homomorphism by
Since the toric divior complement can be identified with a complex torus , we obtain . Since is a complex subtorus in , we obtain the identification . Since we defined the -action on through the map and the -action on (see [Yam24b]), we see that the -action on is effective.
If is a complex submanifold in , then there exists an embedding . Thus, a symplectic form on is , where is a symplectic form on . Let be a map defined by
where is the -action on . By the definition of the -action on , we obtain
for any . Thus, the -action on preserves the symplectic form on . We show that the map is the moment map for the -action on , where is the moment map for the -action on . By the definition of the -action on , the map is equivariant with respect to the -action on . For any point and , we have
From this calculation, we obtain
Therefore, the map is the moment map for the -action on equipped with the symplectic form . ∎
Note that the above discussion is valid for any codimension of .
The vectors used here have the following relation:
Lemma 2.6 ([Yam24b, Lemma 4.14]).
holds.
In this case, the rank of the Jacobian matrix is independent of the choice of of the affine subspace [Yam24a, Proposition 2.12], i.e. we may assume that .
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 be a vertex of .
-
•
If , then for any .
-
•
If , then for any .
Proof.
We assume that . Then, since the defining equation for is
we obtain that for any .
Similarly, if we assume that , then we obtain for any . ∎
Lemma 2.8.
If (or ), then is nonsingular.
Proof.
Assume that . From the proof of Lemma 2.7, the defining equation for is
Since we obtain
the points where the rank of the Jacobian matrix of is equal to zero should be in . However, since from Lemma 2.7, the rank of is equal to one, i.e. does not have a singular point.
Similarly, we assume that , then we obtain that the rank of is equal to one. ∎
Lemma 2.9.
Assume that
or
If , then is nonsingular.
Proof.
Assume that . From the defining equation for , we obtain that
If , then
Therefore, the rank of the Jacobian matrix is equal to one, i.e. is nonsingular. ∎
3. Delzant Polytopes of Torus-equivariantly Embedded Toric Hypersurfaces
In this section, we show the first main result (Theorem 3.15).
Since is surjective, we have the following:
Lemma 3.1.
Let be linearly independent. There exist such that are linearly independent.
Hereafter, we assume that otherwise specified.
3.1. Case of
Let be a vertex of the Delzant polytope . Define . In this section, we show that if the vertex of does not satisfy the condition (2) in Definition 1.1, then is nonsingular.
Lemma 3.2.
The vertex of does not satisfy the condition (2) in Definition 1.1 if and only if .
Proof.
If the vertex of does not satisfy the condition (2) in Definition 1.1, then there exists a direction vector such that the vector is zero. Note that the pullback is given by
Since , we obtain , i.e. . In particular, .
If , then there exists an element . By the definition of the set , we have , i.e. the vector is zero. ∎
In particular, is equivalent to .
Lemma 3.3.
If , then or .
Proof.
Let . By the definition of , we obtain . Since the vectors are nonzero, there exists such that . Since (see [Yam24b, Lemma 3.10]), we have
for any . If , then for any , i.e. . If , then for any , i.e. . ∎
Corollary 3.4.
If , then is nonsingular.
For a vertex of , we define the cone .
Proposition 3.5.
If , then is a vertex of the convex polytope .
Proof.
From Lemma 3.1, we may assume that the vectors are linearly independent. Then, , i.e. .
If there does not exist any nontrivial subspace in the cone , then is a vertex in the polytope . Let be a subspace contained in the cone . For any element , there exist such that
Since is a linear space, . There exist such that
Let for . Since , we obtain
Since we assume that are linearly independent, we obtain that . Since , we obtain , which means that . Therefore, we obtain , i.e. there is no nontrivial subspace in the cone . The point is a vertex of the convex polytope . ∎
This proposition tells us that if the vertex of does not satisfy the condition (2) in Definition 1.1, then the vertex satisfies the condition (1).
Proposition 3.6.
If , then the vectors for are the direction vectors from the vertex of the polytope .
Proof.
Let , i.e. . From Lemma 3.1, the vectors are linearly independent.
For the rest of the proof, we show that
Since it is clear that the right hand side is a subset of the left hand side, we show the other inclusion. For any , we obtain
Therefore, are the direction vectors from the vertex . ∎
3.2. Case of is not a vertex
Let be a vertex of the Delzant polytope . In this section, we show that if the vertex of does not satisfy the condition (1) in Definition 1.1, i.e. there is a nontrivial subspace in the cone , then is nonsingular. From Proposition 3.5, this vertex satisfies the condition (2), i.e. .
Let be a nonzero element in the nontrivial subspace . Since , we obtain . Since the subspace is in the cone , there exist such that
Let for . Then, we obtain
(3.1) |
From Lemma 3.1, we may assume that the vectors are linearly independent.
Lemma 3.7.
Assume that the vectors are linearly independent. If is not a vertex of the polytope , then in Equation 3.1.
Proof.
Assume on the contrary that . Then, from Equation 3.1 we obtain
Since the vectors are linearly independent, we obtain . Since and , we obtain for any , i.e. . This is contradiction to the assumption that is a nonzero element in the subspace . ∎
Lemma 3.8.
Assume that the vectors are linearly independent. If is not a vertex of the polytope , then .
Proof.
From Lemma 3.7, . Then, Equation 3.1 follows
Since from Lemma 2.6 we calculate
we obtain
Since we assume that the vectors are linearly independent, we obtain
(3.2) |
Since means that , we say that
(3.3) |
Assume on the contrary that . Then, from Equation 3.2, we obtain . Since form a -basis of , if , then . This is contradiction to the assumption that is a basis of the orthogonal subspace to the -dimensional subspace in . Therefore, . ∎
Lemma 3.9.
If is not a vertex of the polytope , then or .
Proof.
From Lemma 3.1, we may assume that the vectors are linearly independent.
Corollary 3.10.
If is not a vertex of the polytope , then is nonsingular.
3.3. Case of and is a vertex
Let be a vertex of the Delzant polytope . In this section, we consider the case when the vertex of is a good vertex with respect to the map in the sense of Definition 1.1. In this case, we suppose the following:
Assumption 3.11.
The vectors are linearly independent. Moreover, we have
Assumption 3.11 implies that is a vertex of the polytope . Note that if the polytope is simple, then Assumption 3.11 holds.
Lemma 3.12.
Under Assumption 3.11, if the vertex of is a good vertex with respect to the map , then we obtain
or
Proof.
From Assumption 3.11, there exist such that
(3.4) |
From Lemma 2.6, we calculate
Since are linearly independent, we obtain
(3.5) |
Note that if , then .
Assume on the contrary that . If , then we obtain
(3.6) |
from Equation 3.5. Since the vectors form a -basis of , Equation 3.6 means that , i.e. . This is contradiction to the assumption that is the orthogonal subspace to the -dimensional subspace in . Therefore, .
Since , we obtain , i.e. or . From Equation 3.5, if , then . Similarly, if , then . ∎
By the definition of the pullback , it is clear that the vectors are integral vectors.
Lemma 3.13.
Under Assumption 3.11, if the vectors form a -basis of , then .
Proof.
Corollary 3.14.
Assume that the vertex of the Delzant polytope is a good vertex with respect to the map . Under Assumption 3.11, if there exist distinct such that the vectors form a -basis of , then 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 is good with respect to the map , then the rank of the Jacobian matrix is equal to one at any point for any vertex of . In particular, is a smooth complex hypersurface in the toric manifold corresponding to .
Proof.
It is sufficient to consider each vertex of the Delznt polytope .
If a vertex of is not a good vertex with respect to the map , then the vertex of does not satisfy the condition (1) or (2) in Definition 1.1. If does not satisfy the condition (2), i.e. , then Corollary 3.4 tells us that is nonsingular. If does not satisfy the condition (1), then Corollary 3.10 tells us that is nonsingular. Thus, if a vertex of is not a good vertex with respect to the map , then is nonsingular.
Since we assume that the Delzant polytope is good with respect to the map , if a vertex of is a good vertex with respect to the map , then Assumption 3.11 holds for such and there exist distinct such that the vectors form a -basis of . In Corollary 3.14, we show that is nonsingular in this case.
Therefore, if is good with respect to the map , then is nonsingular for any vertex of , i.e. is a smooth complex hypersurface in the toric manifold . ∎
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 is equal to one at any point of the zero locus of for any vertex of the Delzant polytope, i.e. is nonsingular. In this case, as we show in Proposition 2.5, is toric with respect to the Hamiltonian -action defined in [Yam24b] with the moment map . Moreover, we have the following:
Theorem 4.1 ([Yam24b, Theorem 4.20]).
Assume that is a complex submanifold in the toric manifold . Then, we obtain
where is the Delzant polytope of the toric manifold .
Recall that if is a complex submanifold, then is a symplectic toric manifold with respect to the -action on (Proposition 2.5). By the Delzant correspondence, the moment polytope of the submanifold should be Delzant, i.e. we obtain the following:
Corollary 4.2.
Assume that is a complex submanifold in the toric manifold . Then, the polytope is a Delzant polytope.
In particular, the polytope is simple, i.e. Assumption 3.11 holds under the assumption that is a smooth complex hypersurface.
To show the second main result, we have to consider the case when the vertex of is good with respect to the map .
Hereafter, we assume that the vertex of is good with respect to the map . If is a complex hypersurface, then from Lemma 3.12, we may further assume that and under the Assumption 3.11.
Lemma 4.3.
Assume that the vertex of is good with respect to the map . Under the Assumption 3.11, we obtain and .
Proof.
From Lemma 3.12, we may assume that and .
Assume on the contrary that , i.e. . From Lemma 2.6, we obtain
Since , we obtain , which is contradiction to the assumption that the vertex of is good with respect to the map . ∎
Lemma 4.4.
Assume that the vertex of is good with respect to the map and and . If is nonsingular, then we obtain at least one of the following:
-
(1)
there exists such that and for any ,
-
(2)
.
Proof.
From the defining equation for , we obtain that
Note that and from Lemma 4.3. Assume that . From this equation, we obtain
If is nonsingular, then there exists such that
at the origin . If such , then we obtain for , i.e. and for any . If such , i.e. , then . ∎
Note that we obtain the similar result to Lemma 4.4 when we assume that and .
4.1. The first case in Lemma 4.4
We consider the first case in Lemma 4.4. Hereafter, we assume that and for simplicity when we consider the first case in Lemma 4.4.
Lemma 4.5.
Assume that the vertex of is good with respect to the map . Moreover, assume that and . Then, we obtain
(4.1) |
Proof.
Remark 4.6.
Lemma 4.7.
Assume the same as in Lemma 4.5. Then, we obtain
Proof.
Since from Lemma 4.5, we obtain
we obtain by using the cofactor expansion that
Since we assume and , we obtain
Thus, we obtain
Since form a basis of , , and
∎
4.2. The second case in Lemma 4.4
Next, we consider the second case in Lemma 4.4, i.e. we assume that and .
Lemma 4.8.
Assume that the vertex of is good with respect to the map and that and . Then, we obtain
(4.2) |
Lemma 4.9.
Assume that and . Then, we obtain
Proof.
Since we see
we obtain by using the cofactor expansion that
Since we assume , we obtain
Thus, we obtain
Since form a basis of , , and
∎
4.3. The Prood of the second Main Theorem
By calculating the determinant of the Gram matrix of , we obtain the following:
Lemma 4.10.
For and , we obtain
Proof.
Since the Gram matrix of is
we obtain
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 form a -basis of . Let be a primitive basis of the orthogonal subspace to in . For and , we obtain
Proof.
By the assumption, there exists a vector such that form a -basis of . Since we have
we obtain by the cofactor expansion that
Since form a -basis of , we obtain . From Lemma 4.10, we obtain
(4.3) |
Since we have
and form a -basis of , we obtain
Since we assume that the vector is primitive, we obtain . Therefore, from Equation 4.3 we obtain the desired equation. ∎
Corollary 4.12.
Assume that form a -basis of . Assume that the vector is a primitive basis of the orthogonal subspace to in . For and , we obtain
We use this fact to show that the second main result:
Theorem 4.13.
If the rank of the Jacobian matrix is equal to one at any point of the zero locus of for any vertex of the Delzant polytope , i.e. is nonsingular, then is good with respect to the map .
Proof.
It is sufficient to consider each vertex of the Delznt polytope .
From Lemma 3.12 and Lemma 4.4, if the vertex of is good with respect to the map , then we obtain the two cases:
-
(1)
there exists such that and for any ,
-
(2)
.
We first consider the first case. In this case, we can assume further that and . Under this assumption, Remark 4.6 shows that the vectors are linearly independent and
From Lemma 4.7 and Corollary 4.12, we obtain
which shows that the vectors form a -basis of .
We next consider the second case. In this case, from Lemma 4.9 and Corollary 4.12, we obtain
which shows that the vectors form a -basis of .
Therefore, is good with respect to the map . ∎
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 -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.