Spherical amoebae and a spherical logarithm map
Abstract.
Let be a connected reductive algebraic group over with a maximal compact subgroup . Let be a (quasi-affine) spherical homogeneous space. In the first part of the paper, following Akhiezer’s definition of spherical functions, we introduce a -invariant map which depends on a choice of a finite set of dominant weights and . We call a spherical logarithm map. We show that when generates the highest weight monoid of , the image of the spherical logarithm map parametrizes -orbits in . This idea of using the spherical functions to understand the geometry of the space of -orbits in can be viewed as a generalization of the classical Cartan decomposition. In the second part of the paper, we define the spherical amoeba (depending on and ) of a subvariety of as , and we ask for conditions under which the image of a subvariety under converges, as , in the sense of Kuratowski to its spherical tropicalization as defined by Tevelev and Vogiannou. We prove a partial result toward answering this question, which shows in particular that the valuation cone is always contained in the Kuratowski limit of the spherical amoebae of . We also show that the limit of the spherical amoebae of is equal to its valuation cone in a number of interesting examples, including when is horospherical, and in the case when is the space of hyperbolic triangles.
Comments are welcomed.
Introduction
Let be an algebraic subvariety of the -dimensional complex algebraic torus. The (classical) amoebae of , for a varying real parameter , have been much studied in the past few decades in several contexts, including in tropical and non-Archimedean geometry. The amoeba is defined to be the image of under the logarithm map defined for fixed by
Note that the map is invariant under the action of the compact torus . In fact, with respect to the isomorphism where between and , the logarithm map can be seen to be the projection onto the second component(s), followed by the usual logarithm function .
For an algebraic subvariety in , it is well-known that the amoeba stretches to infinity in several directions, also known as its tentacles. These tentacles indicate the directions in which approaches infinity, or in other words, they parametrize the orbits at infinity in its tropical compactification. The precise statement is as follows (cf. [Jonsson]): as , converges to the tropical variety as subsets in , where convergence is defined here in the sense of Kuratowksi (see Definition 3.3). The theory of amoebae and tropicalization connects classical complex geometry with tropical geometry, which studies piecewise-linear objects in Euclidean space, and the study of these interactions is an active research area.
In this paper, we take initial steps toward generalizing the above to the non-abelian setting. More precisely, we choose to view the complex torus as a special case of a spherical homogeneous space in which the group is abelian. We then see whether, or how, the abelian story (as briefly outlined above) might be generalized for non-abelian spherical homogeneous spaces. Before further details, we first briefly recall the setting. Let be a complex connected reductive algebraic group. A homogeneous space of is called spherical if a Borel subgroup of has a dense orbit, with respect to the usual action by left multiplication. (See Section 1 for details.) Thus, the complex torus is the special case in which is the torus itself, is the trivial subgroup, and the Borel subgroup is again the torus itself.
One of our fundamental motivations for this paper is the desire to better understand the -orbit space of a spherical homogeneous space as above, where is a maximal compact subgroup of . Indeed, in the abelian case, the logarithm map can be viewed as the quotient map of by its maximal compact subgroup , and we can view as the parameter space of the -orbits in . Similarly we may ask whether there exists a non-abelian, or spherical, analogue of the classical logarithm map on which allows us to naturally parametrize the -orbit space of . This problem can be viewed as a generalization of the classical Cartan and polar decompositions in Lie theory. However, as far as we know, there is no general and concrete description of the -orbit space known. (For some important results in this direction see [Knopetal].) A main motivation for the present paper is the belief that there should be a natural isomorphism between and the real closure of the valuation cone (see Section 1.1 for the definition) of an arbitrary spherical variety (over ). In the abelian case, the valuation cone is precisely , so the logarithm provides this natural isomorphism. We think of this paper as taking some steps forward in the search of such a natural isomorphism (see also Conjecture 6). (We also note that related ideas have appeared in [Feu].) However, it should be noted that the results we obtain are of a different nature than a direct generalization of the torus case. In contrast to the abelian situation in which is an isomorphism , what we see in our results is that, in the limit as , the image of our spherical logarithm map (to be defined precisely below) converges in an appropriate sense to the valuation cone. In particular, for fixed , our does not provide the desired canonical isomorphism.
In addition to the above, we were motivated by the notion of a spherical tropicalization map associated to a spherical homogeneous space . Spherical tropical geometry was introduced and developed by Tevelev and Vogiannou [Vogiannou] and by the fourth author and Manon [Kaveh-Manon]. We believe that there should be a theory of spherical amoebae for which correspond to the classical amoebae in the abelian case, such that the “spherical amoebae converge to the spherical tropicalization” in an appropriate sense (see also Remark 3.2).
With these motivations in mind, we now describe more precisely the contributions of this article. Let be a spherical homogeneous space as above. For technical reasons we additionally assume that is quasi-affine. (This is not terribly restrictive, as we explain in Remark 1.6.) Let denote the semigroup of -highest-weights of irreducible -representations appearing in the coordinate ring , considered as a -module in the natural way. Following some rather little-known work of Akhiezer [Akhiezer-rk], we define a -invariant spherical function on associated to a -weight for . We normalize these spherical functions by setting . Our first observations concerning Akhiezer’s spherical functions is the following; see Propositions 2.6 and 2.11 for precise statements.
Theorem 1.
Suppose generate as a semigroup. Then the corresponding spherical functions generate the algebra of -invariant functions on . It follows that the image of the spherical logarithm map corresponding the choice of , parametrizes the -orbit space . Moreover, the spherical functions separate -orbits. More precisely, for any two distinct (hence disjoint) orbits -orbits and in , there exists some such that .
Motivated by the above, let . We define by:
(details are in Section 2). For a fixed , we then define the spherical logarithm map (depending on and with base ) to be the map (see Definition 2.12). 111Note that the spherical tropicalization map defined by the fourth author and Manon in [Kaveh-Manon] naturally takes values in a -vector space. However, since our spherical logarithm map uses a logarithm, we need to work with coefficients. If we assume that the above subset generates as a semigroup, then from Theorem 1 it follows that we obtain a parametrization of , i.e., the points in the image parameterize -orbits in . This is a positive step in the broader program of understanding . We will have more to say about the relation between the image of and the valuation cone below.
The spherical functions used to define have somewhat subtle properties. The next result, below, gives an inequality involving a product of two spherical functions. The essential point here is that spherical functions are not necessarily multiplicative: given two -weights and , in general we can have (see Example 2.10). However, it is worth emphasizing that when is horospherical the multiplicativity indeed holds (Lemma 2.5). When is not horospherical, our next result gives some information about the relation between the product and the other spherical functions. To state it, we need the following notation. For a -weight , let denote the irreducible -representation with highest weight . Given a -module we say appears in if the decomposition of into irreducible -representations contains a copy of . Here we view as a -module. Given and both appearing in , we let denote the -span of the set of products where and . We have the following; for a precise statement see Proposition 2.9.
Theorem 2.
Suppose and appear in and suppose appears in . Then there exists a constant such that for all we have
In particular, since always appears in , we see that there is a constant such that for all .
As we have observed, our definition of depends on the choice of . In the special case when is horospherical and thus the spherical functions are multiplicative, we can define the spherical logarithm map canonically (i.e. independently of the choice of ) by
where is defined by , . In this case, the image of the map lands naturally in the -dimensional vector space , where is the rank of the spherical variety . This is also the vector space in which the valuation cone naturally lies (see Section 1, as well as Remark 2.13). As we mentioned above, this paper was partly motivated by the search for a natural isomorphism between and the valuation cone. Since parametrizes , it is natural to ask for a relationship between it and . A choice of generating set gives a linear embedding of the valuation cone into , the codomain of in the general case, given by:
When comparing with in what follows, we identify with its image in under the above embedding.
We now turn to the related question of spherical amoebae. Let be a subvariety of . We define the spherical amoeba of to be the image of under the spherical logarithm. One of our motivations for this manuscript was to ask: do the spherical amoebae approach (in a suitable sense) the spherical tropicalization of as ? Of course, as in the classical (abelian) case, one needs to be precise about what the word “approach” means. Following [Jonsson], here we take it to mean the convergence of subsets in the sense of Kuratowski (Definition 3.3). Moreover, in the non-abelian case, the image of the spherical homogeneous space under the spherical tropicalization map is not necessarily the entire codomain; instead, the image is the valuation cone . This is in contrast to the classical case , in which the image is the entire vector space . We ask the following (cf. Question 3.4):
(1) Under what conditions do the images approach the valuation cone as in the sense of Kuratowski convergence of subsets? (2) For a subvariety, under what conditions do the spherical amoebae approach as in the sense of Kuratowski convergence of subsets? |
In Remark 3.6 we sketch a proof strategy to give a positive answer to Part (1) of the above question; moreover, as we discuss further below, in Section 4 we show that in many examples, the Kuratowski limit of is indeed the valuation cone . We expect that a proof of Part (2) to be more difficult, and (as in the torus case [Jonsson]), an answer might require tools from non-Archimedean geometry. Nevertheless we conjecture the following.
Conjecture 3.
If is horospherical, and is a subvariety, then the statement in the Question (2) (as stated above) holds, i.e., the spherical amoebae approaches in the sense of Kuratowski as .
We also note that progress on (1) above would partially answer the broad motivating question of finding a relation between and . In this paper, we prove some preliminary results which address the question (1) above. Firstly, we answer the “curve case”, as follows. The precise statement is Theorem 3.1.
Theorem 4.
Let be the -invariant valuation associated to a formal Laurent curve in , convergent for small enough . Let and let be the corresponding spherical function. Then for any highest weight vector we have
Corollary 5.
(1) The Kuratowski limit, as , of the image contains the valuation cone . (2) When is horospherical, this limit is the entire vector space, which in this case coincides with the valuation cone.
Statement (1) in Corollary 5 shows that is contained in the Kuratowski limit of the images . When is the limit exactly ? In Section 4, we give some examples for which, with appropriate choices of , the images limits to as : (i) the basic (quasi) affine space (cf. Example 1.1(4)), (ii) the “group case” with acting on (cf. Example 1.1(5)), (iii) , and (iv) the space of hyperbolic triangles. These examples give some evidence that the answer to the question (1) above may be positive, and hence the -orbits is related, in an appropriate limit, to . In fact, the first author previously conjectured a tight relation between and . Since the conjecture is not recorded elsewhere, we put it here.
Conjecture 6 (Batyrev).
The -orbit space is a stratified manifold with corners where the boundary faces are in natural bijection with the faces of the valuation cone , and the stabilizers of the points in the relative interior of are maximal compact subgroups in the corresponding satellite subgroup of (as defined in [Batyrev-Moreau]). Moreover, is homeomorphic to the valuation cone as a stratified space, i.e., there is a homeomorphism that, for every face , restricts to a homeomorphism between the relative interior of and that of .
As a special case we also state the following conjecture.
Conjecture 7.
A spherical homogeneous space is horospherical if and only if the -orbits space is homeomorphic to a Euclidean space.
We believe that the spherical logarithm maps may provide a tool to attack Conjecture 7.
For historical context, we should note here that it has been known for some time that the theory of moment maps from symplectic geometry can be used to parametrize -orbits for multiplicity-free spaces. The theory of multiplicity-free Hamiltonian spaces are related to this question of -orbits on spherical for the following reason: if is smooth spherical variety embedded in a projective space which is equipped with a linear -action and a -invariant Hermitian structure (and hence has a -invariant Kähler structure), then is a multiplicity-free Hamiltonian -space. Thus, a -equivariant embedding of into a projective space gives such a structure. Moreover, for a multiplicity-free Hamiltonian -space , it is known that the Kirwan map222 The Kirwan map is the composition of the -moment map with the quotient map where denotes the Lie algebra of the (compact) maximal torus of . provides a parametrization of the -orbits in (see [Brion-moment, Proposition 5.1]). Thus, this Kirwan map provides another potential approach to Conjectures 6 and 7.
We now outline the contents of the paper. In Section 1 we briefly recall necessary background, including the definition of the spherical tropicalization map as in [Vogiannou, Kaveh-Manon]. Then in Section 2 we define the spherical functions, following Akhiezer, and prove several properties about them. We also define the notion of a spherical logarithm map. In Section 3 we give precise statements of the Kuratowski convergence of subsets, and formulate our questions concerning the convergence of spherical amoebae. We also prove the “curve case” recounted above, as well as the horospherical case. Finally, in Section 4 we compute spherical functions in several important examples, and in some examples, we also give some natural parametrizations of the -orbit space .
Acknowledgements. We would like to thank Dmitri Timashev for invaluably helpful personal discussions and correspondences clarifying several details. In particular, Example 2.10(1) is due to him. We also thank Stéphanie Cupit-Foutou for helpful correspondence (and in particular for Remark 2.8). Some of this work was conducted at the Fields Institute for Research in the Mathematical Sciences during the fourth author’s stay there as a Fields Research Fellow, and we thank the Fields for its support and hospitality. The second author was supported in part by a Natural Science and Engineering Research Council of Canada Discovery Grant and a Canada Research Chair (Tier 2) from the Government of Canada. The fourth author was partially supported by a National Science Foundation Grant (DMS-1601303) and a Simons Collaboration Grant for Mathematicians.
1. Preliminaries on spherical varieties and tropicalization
In this section, we set some notation and briefly review some facts regarding spherical varieties.
1.1. Background on spherical geometry
Let be a connected reductive algebraic group over . Let denote a choice of Borel subgroup of and let be a maximal torus with . The weight lattice of is denoted by , and the semigroup of dominant weights corresponding to the choice of is correspondingly denoted by . The cone generated by is the positive Weyl chamber . For a dominant weight we denote the irreducible -module with highest weight by . We usually denote a highest weight vector in by .
Recall that a normal -variety is called spherical, or a spherical (-)variety, if there exists a dense -orbit in . (Since all the Borel subgroups are conjugate, this condition is independent of the choice of Borel subgroup ). A homogeneous space is called spherical if it is spherical with respect to the left action of on itself by multiplication.
Let be a spherical -variety and let be the field of rational functions on . Since acts on , there is also a natural -action on . We say that is a -eigenfunction (sometimes also called -semi-invariants) if it has the property that for some -weight where denotes the character associated to , i.e., for all we have . Let denote the set of -eigenfunctions in , and let denote the subset of consisting of all weights arising from -eigenfunctions, namely,
Consider the map associating to a -eigenfunction its corresponding weight (in the notation above, the map sends to ). It is clear that this is a group homomorphism with respect to multiplication in and addition in and that its image is by definition. Hence is a sublattice of . Moreover, since is spherical and thus there exists an open dense -orbit, it follows that this map has kernel precisely the (non-zero) constant functions , thus inducing an isomorphism between and .
We briefly recall some standard examples of spherical varieties.
Example 1.1.
-
(1)
Let be an algebraic torus. In this case, , and therefore a spherical -variety is the same as a toric -variety.
-
(2)
Let be a partial flag variety, equipped with the usual left action of by multiplication. By the Bruhat decomposition, is then a spherical -variety. This is an example of a projective spherical -variety for a non-abelian .
-
(3)
Let and cnsider the natural action of on by the usual (left) matrix multiplication. It is straightforward to see that acts transitively on . Moreover, the stabilizer of the point is the maximal unipotent subgroup of upper triangular matrices in with ’s on the diagonal. Thus can be identified with the homogeneous space . Let be the subgroup of upper triangular matrices in . Then it is not hard to see that the -orbit of the point is the dense open subset of . Thus, is a spherical variety. This is an example of a quasi-affine (but not affine) spherical variety. Similarly, it can be verified that is a spherical variety for the natural action of . (However, for , the -stabilizer of a point in is larger than a maximal unipotent subgroup).
-
(4)
More generally, consider where is a maximal unipotent subgroup of , equipped with the left action of . Again by the Bruhat decomposition, is a spherical -variety, and it is well-known that is quasi-affine. It is useful to note that there is a natural fiber bundle , where is the Borel subgroup containing , with fiber isomorphic to the torus .
-
(5)
Let and consider the “left-right” action of on given by . Clearly this action is transitive, and the stabilizer of the identity is the diagonal subgroup . Thus can be identified with the homogeneous space . Again from the Bruhat decomposition it follows that is a -spherical variety. Here the Borel subgroup of is chosen to be , where is a Borel subgroup of and is its opposite. The -equivariant completions of are usually called group compactifications.
For the rest of the paper we restrict attention to the setting of spherical homogeneous spaces, i.e., . In the setting of spherical homogeneous spaces , it turns out there is a convenient way to view the sublattice , as follows. Choose a Borel subgroup so that the -orbit of is dense in ; such a exists since is spherical. Now let be a maximal torus in the intersection and choose to be a maximal torus in containing . Define . It is known that the character lattice of can be identified with .
We next briefly review the theory of valuations on . In this manuscript, by a valuation on we will mean a discrete -valued valuation on , i.e., we assume
-
(1)
and ,
-
(2)
or ,
-
(3)
,
-
(4)
,
-
(5)
.
A valuation is -invariant if for any and we have . We define
(1) |
From the Luna-Vust theory, it is known that any -invariant valuation in the sense explained above is a geometric valuation, or more precisely, is induced by (up to multiplying by a constant in ) a divisor (i.e. is the order of vanishing along a divisor).
Example 1.2.
Let and . Then . For any vector we can construct a -invariant valuation as follows. For we define
Then extends to the field of rational functions and this is easily checked to be -invariant.
Let be a -invariant valuation. By the definition of valuations, gives rise to a homomorphism , where the group operation on is multiplication of functions. Moreover, by assumption, the valuation evaluates trivially on constant functions, so in fact we obtain a group homomorphism . The identification above then allows us to view this as a linear map . We denote this linear map by . We introduce the notation for this -vector space. With this notation, the correspondence is a mapping
(2) |
The following is well-known [Luna-Vust, 7.4 Prop].
Theorem 1.3.
With the assumptions as above, the map is injective, i.e., a -invariant valuation is uniquely determined by its restriction to the -eigenfunctions.
Due to the above theorem, we may henceforth identify with its image .
Note there is a natural pairing between -invariant valuations and the lattice given by .
Example 1.4.
Continuing Example 1.2, the lattice in the case is isomorphic to . The -semi-invariant functions are the monomials , and the monomial corresponds to the -weight . In this special case, it is straightforward to explicitly compute the natural pairing described above. For a -invariant valuation as constructed in Example 1.2, the element sends to (by definition of ), i.e., the usual dot product of with .
The following result is due to Brion [VUGen] and Knop [WeylMom].
Theorem 1.5.
The set is a co-simplicial cone in the vector space . Moreover, it is the fundamental domain for the action of a Weyl group of a root system. More precisely, there exists a set of simple roots in this root system such that the cone is defined by:
where the pairing is the one described above, and the lie in .
The set of simple roots is called the system of spherical roots of . This Weyl group of the spherical root system is also called the little Weyl group of .
Henceforth, we additionally assume that the spherical variety is quasi-affine.
Remark 1.6.
This assumption is not severe. Following [Akhiezer-rk] one can show that, by adding an extra component if necessary, we can assume that is quasi-affine. More precisely, let . Take a character and define . The homogeneous space is spherical for the action of . In fact, if is a Borel such that is open in then is open in where . Also, one can show that for a suitable choice of a character , the homogeneous space can be equivariantly embedded in some finite dimensional -module [Humphreys, Section 11.2] and is thus quasi-affine. The projection , , induces a -equivariant morphism which is a fiber bundle with fibers isomorphic to .
Assuming that is quasi-affine, let denote the ring of regular functions on . When is considered as a -representation in the natural manner, it is known that can be decomposed into finite-dimensional irreducible -representations (see e.g. [Timashev, Appendix p.250]) so we have
(3) |
where denotes the -isotypic component of . When is quasi-affine, sphericity of is equivalent to the ring of regular functions being a multiplicity-free -module [Timashev, Theorem 25.1(MF5)]. The decomposition above motivates the following definition.
Definition 1.7.
We denote by the sublattice of generated by . There is an analogue of the usual dominant order for the sublattice , defined as follows. Let . We say that if is a linear combination of the spherical roots with nonnegative coefficients. We call the spherical dominant order. One has the following [Knop-LV, Section 5].
Theorem 1.8.
Let be a quasi-affine spherical homogeneous space. Let be the ring of regular functions on . Let , . Then the product lies in
Let and let appear in the product , where here denotes the span of all products of the form for . For such a triple , we call the weight a tail. The tail cone of is defined to be the closure of the cone in generated by all the tails. We have the following (see e.g. [Knop-LV, Lemma 5.1]).
Proposition 1.9.
Let be a quasi-affine spherical homogeneous space. Then the tail cone is the dual cone to the (negative of the) valuation cone, .
1.2. Spherical tropicalization
We start by recalling the notions of a germ of a curve and a formal curve (see for example [Timashev, Section 24]). We let denote the algebra of formal power series with coefficients in and its field of fractions, i.e. the field of formal Laurent series with finitely many negative exponents. If we denote by the order of in the Laurent series . Clearly is a -valued valuation on the field .
Let be a variety. A formal curve on is a -point of . An -point on is called a convergent formal curve. The limit of a convergent formal curve is the point on obtained by setting in . If we assume is embedded in an affine space then a formal curve on is an -tuple of Laurent series satisfying the defining equations of in . If is convergent then its coordinates are power series and their constant terms are the coordinates of the limit point .
Definition 1.11 (Valuation associated to a formal curve).
A formal curve on defines a valuation as follows.
(4) |
We recall that the algebraic closure of the field is the field of formal Puiseux series with coefficients in , the elements of which are formal series
where the are non-zero complex numbers for all , and are rational numbers that have a common denominator. (Sometimes is denoted as in the literature.) We call a point in a formal Puiseux curve or simply a Puiseux curve on . Definition 1.11 extends naturally to Puiseux curves. That is, a formal Puiseux curve on gives a valuation , defined by the same equation (4).
Now we restrict attention to the case of spherical varieties and -invariant valuations. Following the setting in the previous section, we assume is a quasi-affine spherical homogeneous space. The main ingredient in the definition of spherical tropicalization, Definition 1.13 below, is the construction of a -invariant valuation from a given arbitrary valuation on . The following well-known result is key (see [Knop-LV, Lemma 1.4], [Sumihiro, Lemma 10 and 11], [Luna-Vust, 3.2 Lemme]).
Theorem 1.12 (Sumihiro).
Let be a quasi-affine spherical homogeneous space. Let be a valuation.
-
(1)
For every , there exists a nonempty Zariski-open subset such that the value is the same for all . We denote this value by , i.e.
-
(2)
For defined as above, we have .
-
(3)
is a -invariant valuation on .
Recall that a formal curve on gives rise to a valuation . We let denote the -invariant valuation constructed by Theorem 1.12 from . Concretely,
for every and is the dense open set in given in Theorem 1.12(1). By Theorem 1.12 we know is a -invariant valuation, i.e., .
Definition 1.13.
Let be a quasi-affine spherical homogeneous space. Following [Vogiannou], we call the map
the spherical tropicalization map.
As in Section 1, it follows from Luna-Vust theory that the map is surjective.
Example 1.14.
Let , equipped with the natural action of as in Example 1.1(3). Thus and the algebra of regular functions is the polynomial ring . It is not difficult to see that coincides with the weight lattice of . Indeed, the function on is a -eigenfunction in whose -weight is . Let be a formal curve in , where and are elements of . Let . We compute that . From the construction of the -invariant valuation above, we know for for some Zariski-open subset. It follows that
(5) |
There is another way of understanding the -invariant valuation associated to a formal curve and that is through the generalized Cartan decomposition for spherical varieties. It goes back to Luna and Vust ([Luna-Vust]). A proof of it can also be found in [Gaitsgory-Nadler, Theorem 8.2.9].
Theorem 1.15 (Generalized non-Archimedean Cartan decomposition for spherical varieties over ).
The -orbits in are parameterized by . Here denotes the lattice dual to the -weight lattice , and a cocharacter corresponds to the -orbit through the formal curve .
Thus the valuation can be interpreted as the valuation represented by the point of intersection of the -obit of in and the image of valuation cone (under the exponential map) in .
Example 1.16 (non-Archimedean Cartan decomposition).
As in Example 1.1(5) consider with left-right action of . Theorem 1.15 applied in this case recovers the a non-Archimedean version of the usual Cartan decomposition (see [Iwahori]). With notation as above, it states that:
Here is the cocharacter lattice and is the intersection of with the dual positive Weyl chamber. We regard both as subsets of .
When the above non-Archimedean Cartan decomposition gives the well-known Smith normal form of a matrix (over the field of formal Laurent series which is the field of fractions of the principal ideal domain , the ring of formal power series).
Example 1.17 (Non-Archimedean Iwasawa decomposition).
As in Example 1.1(4) consider the spherical homogeneous space where is a maximal unipotent subgroup of . In this case Theorem 1.15 gives a non-Archimdean version of the Iwasawa decompostion (see [Iwahori]). It states that:
where as in the previous example, is the dual lattice to the weight lattice and we regard it as a subset of .
2. Spherical functions, the spherical logarithm, and separating -orbits
In this section, following Akhiezer [Akhiezer-rk], we define certain -invariant real-valued functions on a spherical homogeneous space. As in Section 1, we assume throughout that is a quasi-affine spherical homogeneous space. Let be a maximal compact subgroup of .
Let be a (non-zero) irreducible -representation in . Fix a -invariant Hermitian product on . Let be an orthonormal basis with respect to this -invariant Hermitian product. Define the function by:
(6) |
Lemma 2.1.
.
Proof.
Suppose . From the definition it follows that for all . Since is a -representation, for any fixed , we know is a linear combination of ’s, but this implies also. Since this was true of any , we conclude that for all . This implies as functions on , for all . This is a contradiction, since is a non-zero representation. ∎
By the above lemma, we may normalize our choice of so that it satisfies
Remark 2.2.
It is not hard to see that, with respect to any -invariant Hermitian product on , two -weight vectors corresponding to different -weights must be orthogonal with respect to . Thus we may assume without loss of generality that the orthonormal basis consists of -weight vectors.
The following lemma is straightforward.
Lemma 2.3.
The function is -invariant, and independent of the chosen orthonormal basis .
We call the function the spherical function associated to the highest weight . Before stating the next result we need some preliminaries. First we consider the algebra of -invariants . By multiplication of functions, we have a natural map where the target denotes the -span of all -valued functions on of the form where . It is possible to see, using the fact that holomorphic functions are determined by their values on any non-empty open neighborhood, and the independence of holomorphic and anti-holomorphic variables and in some local coordinates, that this map is an isomorphism. Therefore, in what follows, we will slightly abuse notation and sometimes write and sometimes write its image . In both cases, the -action is given by a diagonal -action on each factor, and the identification is -equivariant.
The following is a result of Akhiezer [Akhiezer-rk, Lemma 3].
Proposition 2.4.
[Akhiezer-rk, Lemma 3] The set of functions is a basis for as a vector space over . In particular,
(7) |
Proof.
The inclusion “” follows by Lemma 2.3. For the reverse inclusion, first observe that
where denotes the set of -equivariant module homomorphisms and where we have used the fact that, for any rational -module , the complex conjugate -module is isomorphic to the dual -module (the isomorphism depends on the choice of a -invariant Hermitian product on ). By Schur’s lemma, if and otherwise. Thus we conclude that . Since for each we see that the claim will follow if we can show that lies in the image of under the multiplication map and that it is non-zero. But this follows from the definition (6) of and Lemma 2.1. ∎
Multiplication of spherical functions can be understood in terms of the tail cone, as follows.
Lemma 2.5.
Let . Then
for coefficients . Moreover, we have .
Proof.
We follow the proof of Proposition 2.4. We have:
(8) |
where the last inclusion is by Schur’s lemma and the fact that the multiplication map is an isomorphism (so we may identify with ). This proves the first claim. To prove that , we consider the space
(9) |
(so this is the space appearing above, before taking -invariants) as a -representation, for a maximal torus of . From Theorem 1.8 and the definition of a tail, we know that in the usual dominance order for any that appears in the direct sum (9). Since for any weight, denotes the irreducible -representation with highest weight , it follows that the highest -weight that can appear in (9) is (where the partial ordering is if in the usual dominance order). Moreover, from the above it follows that the only summand in (9) which contains as a -weight is , i.e. the term corresponding to .
Recall that the spherical function is a sum (identified with ), where is a basis of -weight vectors in . In particular, has a non-zero component in the one-dimensional -weight space of weight , i.e. the highest-weight space of . It follows from the above discussion that in order to determine whether , it suffices to show that has a non-zero component in the -weight space in (9) of weight . For the following we temporarily denote by (respectively ) the highest weight vector of -weight (respectively ) in (respectively ). By definition, and similarly for . Hence . Since has weight and is non-zero, and all other terms are strictly smaller, there can be no cancellation and we conclude as desired. ∎
Proposition 2.6.
If generate as a semigroup, then generate as an algebra.
Proof.
The proof is a standard “canceling the leading term argument”. First we consider the case where is semi-simple. One can construct a well-ordering on the semigroup respecting the addition, such that if is a non-negative combination of spherical roots then . Take a vector in the interior of the dual cone to the positive Weyl chamber. Moreover, assume that is irrational, that is, for any nonzero weight , . For two weights , define if and only if . It is easy to see that has the required properties.
For let and define
For convenience we will write instead of for . Suppose by contradiction that there exists that cannot be represented as a polynomial in the and is minimum among all such -invariant functions. Then since generate we can find such that . It follows that there is such that . Thus , and hence , can be represented as a polynomial in the . The contradiction proves the claim.
Next suppose is reductive. Replacing with a finite cover if necessary we can assume that where is connected semisimple and is the connected component of the center of . Then as semigroups , where is the character lattice of the torus . One shows from the definitions that if and then
(10) |
The claim follows from (10) and the cancelation of the leading terms argument as in the semisimple case. ∎
Remark 2.7.
Let be a nonzero function. Consider the function obtained by averaging over , i.e.,
where is the Haar measure on . Since (the span of) is a finite-dimensional vector space, the integral is still an element in , and is -invariant by construction. From Proposition 2.4, it follows that is a scalar multiple of .
Remark 2.8 (Highest weight monoid not necessarily free).
In general, the highest weight monoid may not be freely generated as a monoid. Here is an example which was communicated to us by Stéphanie Cupit-Foutou and is inspired by [Luna, §6.3]. Let and for let be a maximal tori of and its normalizer in . Now let and where , are the non-trivial characters of and respectively. Then is spherical and affine, and its weight monoid is not free: it is generated by , and where denotes the fundamental weight for .
Proposition 2.9.
Let and suppose is a tail, i.e. appears in , then there is constant such that for all we have:
Proof.
By definition and where , are orthonormal bases for , respectively. Let be an orthonormal basis for . As , we can write for some . For , it follows that
where , the first inequality is the triangle inequality and the third inequality is the Cauchy-Schwarz inequality. Squaring both sides we obtain where the constant is , so by dividing we obtain the claim. ∎
The following examples show that spherical functions need not be multiplicative. More precisely, for , is not necessarily a scalar multiple of .
Example 2.10.
(1) This example is due to Dmitri Timashev. Let and let and . Then it can be shown that is a spherical variety and can be realized as the affine variety in defined by the equation . (All facts claimed here without proof are shown in Section 4.4.) It turns out that is generated by , corresponding to the standard representation of and its dual. Let and be the corresponding spherical functions. We wish to compare the product with the spherical function corresponding to .
Let and denote the standard basis of and its dual basis of respectively. The action of on the coordinate ring
is the standard one, viewing (respectively ) as the coordinates on (respectively ) corresponding to (respectively ). It is not difficult to see that the span of the (respectively ) yields a representation of isomorphic to , the dual of the standard representation (respectively , the standard representation). The corresponding spherical functions are therefore , where are the coordinates of , and , where are the coordinates of . Hence their product is .
Next we wish to compute the spherical function corresponding to . Since the irreducible -representation of highest weight is the adjoint representation , we can proceed as follows. Suppose there exists a -equivariant embedding . This induces a -equivariant map backwards . (Note that so in our example we may identify with .) Provided that the embedding is nontrivial, by Schur’s Lemma this map must take to the (isomorphic copy of) appearing in the decomposition of into irreducible -modules. The spherical function is then the pullback via the embedding of the sum of the norm-squares of the coordinates on (with respect to some -invariant inner product on ). The Hermitian inner product on is -invariant with respect to the adjoint action of . Next, notice that the map gives a -equivariant morphism from to , where we view as . Thus we obtain
(11) |
which is clearly not proportional to .
(2) (Group case) Consider as a -spherical homogeneous space. Let be a Borel subgroup of and take as Borel subgroup of where denotes the opposite Borel subgroup to . Then the highest weight semigroup of is . For let us write in place of the spherical function . By reasoning similar to that in example (1) above, we have
where is the irreducible representation of with highest weight and stands for the Hermitian adjunction with respect to some -invariant Hermitian product on the representation space . It follows that if lies in the real locus of a split maximal torus , then where denotes the character of the representation . But one knows that the product of characters is the character of tensor product. This shows that the multiplicativity does not hold for the .
Proposition 2.11.
Let be a quasi-affine spherical homogeneous space. Then the spherical functions , , separate -orbits in . More precisely, for any two distinct (hence disjoint) -orbits and in , then there exists some such that .
Proof.
Since is quasi-affine, there exists a -equivariant embedding for some , where is equipped with a linear -action (see [Popov-Vinberg, Section 1.2]). Let be two distinct -orbits. Since is compact, both are compact. As disjoint compact subsets of an affine space, by Urysohn’s lemma, there exists a continuous, real-valued function on such that and . By the Stone approximation theorem (see e.g. [Cheney, Chapter 6]), there exists a real polynomial on that approximates arbitrarily well on any compact subset. Since acts linearly on the ambient , we can average this polynomial by integrating over and obtain a -invariant real polynomial that separates and . The restriction of to belongs to and hence we conclude that there is some that separates and as required. ∎
Let be a set of semigroup generators of . We have seen in Proposition 2.6 that the collection of associated spherical functions are algebra generators of the subalgebra . Thus, it follows from Proposition 2.11 that they also separate -orbits. Thus, this set of spherical functions plays an important role in understanding the geometry of -orbits of , and motivates the following definition of the spherical logarithm map.
Definition 2.12.
Let be a set of semigroup generators of and let denote the corresponding spherical functions. We define
We call this the spherical logarithm map associated to .
Remark 2.13.
We already noted in the introduction that when is horospherical, we can define a spherical logarithm map canonically, independent of a choice of . In fact, we can say more: in the horospherical case, we also do not need to assume that is quasi-affine. Here we briefly sketch how to define the spherical logarithm map without quasi-affineness.
First consider an -dimensional torus with character lattice and maximal compact torus . Then there is a canonical logarithm map defined as follows: for , let be the homomorphism of that sends any character to , where denotes the natural logarithm. For , we then define . It is clear that is -invariant for any .
We now return to the spherical setting. Suppose is horospherical. In this case it is known that is a parabolic subgroup. Moreover, for a Levi decomposition of , we have for some subgroup where and denotes the commutator subgroup of [Timashev, Lemma 7.4]. It follows that the quotient group is abelian and hence a torus. For the natural projection, we then see that all the fibers of are tori isomorphic to , and in particular, . Now let be a maximal compact subgroup of such that . Then acts transitively (from the left) on . Since the projection is -equivariant and acts transitively on the base, it follows that every -orbit in intersects . This yields a map from the space of -orbits to (it turns out the intersection of a -orbit with is a -orbit). We can now define the spherical logarithm map by composing with the defined in the previous paragraph. It can be verified that when is quasi-affine this map coincides with the map defined using spherical functions (cf. the discussion after Theorem 1).
3. Spherical tropicalization, spherical amoebae and the limits of spherical logarithms
The purpose of this section is to examine relationships between the spherical tropicalization map, introduced in Section 1.2, with the spherical logarithm map of Definition 2.12. In order to do so, a slight change in perspective is useful, because the spherical logarithm is inherently a real object (it being a logarithm), i.e. defined over , whereas the valuation cone is naturally defined over . More precisely, for the discussion that follows, we first choose an explicit identification of with using a choice of basis. Specifically, let be the basis of appearing in Definition 2.12. This gives us an identification . Moreover, we additionally extend our coefficents from to for the remainder of this manuscript, so we consider and the choice of basis above naturally also yields an identification , which we fix throughout. The embedding in (2) then allows us to think of the valuation cone as a subset of . We let denote the closure of in with respect to the usual Euclidean topology on . It is a co-simplicial cone in with the same defining equations as given in Theorem 1.5. (We slightly abuse terminology and call also the valuation cone.) The reader may note that the spherical logarithm map of Definition 2.12 uses a choice of where it is possible that . In such a case, the image of lies in and not in , whereas the valuation cone lies in , as just described above. In this situation we will use the linear embedding of the valuation cone into specified by , given by:
In this way we can embed in and identify it with its image in , thus allowing us to compare with the image of the spherical logarithm.
Our first result shows that points in the valuation cone can be realized as limits of points in the image of the spherical logarithm map. As in the previous section, we assume here that is a quasi-affine spherical homogeneous space. Let be the Borel subgroup of such that the -orbit of is open in . Note that the multiplication map (similar to the Iwasawa decomposition) is a surjection, with kernel isomorphic to .
Theorem 3.1.
Let be a quasi-affine spherical homogeneous space. Let be a convergent formal Laurent curve in which is convergent for sufficiently small . Let be the -invariant valuation associated to the formal curve . Let and let the corresponding spherical function. Then for any highest weight vector we have:
(12) |
Proof.
Let be fixed throughout. By Theorem 1.12 there exists an open set such that for we have where computes is the order in of after restricting to the curve . We claim that there is an open dense subset such that for we have
(13) |
To see this, first note that for if we write with and then since is a -weight vector we have . Now let be the multiplication map and the projection on the first factor. It suffices to take . Clearly, is open. Note that the multiplication map is a fibration with fibers isomorphic to . But in a fibration the inverse image of a dense subset is dense. This shows that is dense in as well. Since the projection map is open, it follows that is open and dense in , which proves the claim.
Now let us define the -invariant function by
(14) |
where is the Haar measure on . By Remark 2.7 we know that is a scalar multiple of the spherical function .
Let denote the value of the invariant valuation on . Now we claim that the limit
exists and is nonzero. We have just seen in (13) that for we have
for some constant . Hence , for . From the definition 14 of we have
where , and the term-by-term integration which is implicit in the last equality is justified because, from the compactness of , it follows that for sufficiently small , the series converges uniformly in . Finally, since is a constant multiple of , from the above we have:
where the last equality is the definition of . This finishes the proof. ∎
Remark 3.2.
In [Kaveh-Manon, Conjecture 7.1], the fourth author and Christopher Manon conjecture an analogue of the Cartan decomposition for a spherical homogeneous space over . They also suggest a definition of a spherical logarithm map on whenever such a Cartan decomposition exists. In particular, in [Kaveh-Makhnatch], the singular value decomposition theorem (i.e. the Cartan decomposition for ) is used to propose a definition of a spherical logarithm map on the group , and it is proven that the logarithm of singular values of a curve on approach its invariant factors. This result can be regarded as a special case of Theorem 3.1. It should be noted that the definition of spherical logarithm proposed in [Kaveh-Manon] is a priori different from that given in this paper – for instance, in the “group case” , our definition gives (the logarithms of) symmetric functions in the singular values (cf. Section 4.2), whereas the definition in [Kaveh-Manon] gives (logarithms of) the singular values.
Theorem 3.1 holds for any weight . By restricting attention to it follows that
where we think of the RHS as an element of as explained above. Thus Theorem 3.1 says that, for certain curves , the image of under the spherical tropicalization map can also be realized as the limit of points in the image of the spherical logarithm map. This is reminiscent of the phenomenon seen in the classical case of , recounted in the Introduction, where the amoeba – the image under map of – approaches the image of under the valuation map. Motivated by the classical case and by Theorem 3.1, we formulate below a question relating the spherical analogues of and .
To give precise statements, we need the notion of Kuratowski convergence of subsets of a topological space, which can be thought of as a notion of continuity for the association .
Definition 3.3 (Kuratowski convergence (see e.g. [Kuratowski, Jonsson])).
Suppose and are topological spaces. Let and let be the restriction of the projection to . For let . Let and . We say that converges to in the sense of Kuratowski as converges to if the following conditions hold:
-
(a)
For every there exist neighborhoods of and of such that for all .
-
(b)
For every and any neighborhood of in , there exists a neighborhood of such that for all .
Suppose is a quasi-affine spherical variety and is a subvariety. Following the terminology in the classical case, we define the spherical amoebae to be the images in of under the spherical logarithm maps. We call the spherical tropicalization of , considered as a subset of via the identification with discussed previously. With this terminology in place, we may ask the following.
Question 3.4.
Let be a subvariety as above and be a set of semigroup generators of . Then, as , we ask:
-
(1)
Under what conditions do the sets approach the valuation cone in the sense of Kuratowski?
-
(2)
Assuming that the images converge to in the sense of Kuratowski, under what additional conditions on the subvariety do the spherical amoebae approach in the sense of Kuratowski?
We note here Theorem 3.1 already gives us some information about Question 3.4. The following are straightforward corollaries of Theorem 3.1.
Corollary 3.5.
Let be a quasi-affine spherical homogeneous space. Let be a set of semigroup generators of . Let denote the valuation cone of .
-
(1)
The Kuratowski limit, as , of the images contains the valuation cone.
-
(2)
When is horospherical this limit is the entire vector space, which in this case coincides with the valuation cone.
Proof.
The first claim follows immediately from Theorem 3.1. The second claim uses the fact that, in the horospherical case, the valuation cone is the entire vector space . Since by (1) the limit of the images must contain the valuation cone, this means the limit must be all of . ∎
Although we do not have a complete answer to Question 3.4(1), in the remark below we give a sketch of an argument which may be useful.
Remark 3.6.
It may be possible to answer Question 3.4 (1) positively using Proposition 2.9 as follows. Suppose lies in the Kuratowski limit as of the sets . Then by definition there exist sequences , with , and , , such that . We would like to show that lies in the image of the valuation cone in . By Proposition 2.9 we know there exists a constant such that
Taking of both sides we have:
Now observe that in order to show it would be enough to find such that for any highest weight we have
Then by the above inequalities we have , for all , , such that is a tail and hence . We expect that the proof that such a exists would be a variant of the argument in the proof of Theorem 3.1. The case when , namely when the highest weight monoid is freely generated, might be easier to work out.
In the next section, we show that in a number of interesting cases, the sets not only contain the valuation cone, but in fact they do limit to exactly the valuation cone in the Kuratowski sense.
4. Examples
In the previous section, we asked in Question 3.4 for conditions under which, in analogy with the classical case, the spherical amoebae approach the spherical tropicalization. In this section we focus on Question 3.4(1), i.e. the case when we take . We analyze the following interesting cases of spherical homogeneous spaces: (1) the “group case” of acting on , for , (2) the basic affine space , (3) the case , and (4) the case of the “space of hyperbolic triangles” (the precise definition is in Section LABEL:subsec:_hyperbolic_triangles). In the cases of the basic affine space and , it turns out that the spherical amoebae coincide with the valuation cone for all , so the limit does indeed coincide with the valuation cone, with a rather trivial limiting process. However, for the “group case” and the hyperbolic triangles case, the spherical amoeba is different from the valuation cone but the limit does indeed approach the valuation cone via a non-trivial limiting process (see Figure 1).
As in Section 3, throughout this section we work over .
4.1. Two small examples with
4.1.1. The case of
Let and let , the maximal (diagonal) torus of . Consider the action of on by left multiplication, and the corresponding diagonal action of on . Then it is not hard to see that the stabilizer of the point is precisely , and that the orbit of under the -action is , where denotes the diagonal copy of in the direct product. The compact group is in this case. The (diagonal) action of on has moment map
where are homogeneous coordinates on (so are considered as elements of ) and denotes the identity matrix. Composing this with the map which quotients by the coadjoint action of on , i.e. the so-called “sweeping map” , yields that the Kirwan map is given by
The Kirwan polytope (i.e. the image of under the above map) is the interval and the image of the diagonal is straightforwardly computed to be , so the distinguished -orbit (the diagonal) corresponds to the boundary value.
There is also another natural parametrization of -orbits which can be described in terms of the angle between two complex lines in . More specifically, given where , we may define
where is the standard Hermitian product on . Geometrically, is the quantity where is the angle between the two complex lines spanned by and , or equivalently, the spherical distance between two distinct points if we represent , by complex vectors with . It turns out that if and only if the two pairs are in the same -orbit, so the function provides a parametrization of -orbits in . It follows from the Cauchy-Schwarz inequality that the image of is the interval . Thus, composing with , i.e. , we obtain an identification of the -orbit space with the valuation cone (which in this case is ).
Finally, there is also our spherical function corresponding to the irreducible representation of of highest weight in . This can be computed to be
One can show that . On the other hand, . This shows that the image of is . Hence we have a parametrization of the -orbits by the points in this interval. One can obtain a parametrization of the -orbit space by (the valuation cone) by taking the limit, as , of the image of .
The above computations show that, in this case, we have three parametrizations of the space of -orbits by (half open) intervals in , namely: the spherical function above, the Kirwan map computed above, and the map as above.
4.1.2. The case of
As in the last section, we take but this time we take . In this situation we have so there is a natural map . The homogeneous space can be identified with where is a smooth conic, as follows: consider the map given by multiplication. Note that and . This product map induces a morphism , where is the smooth conic defined by the vanishing of the discriminant on . One sees that the natural projection is then identified with . The non-identity element in the quotient corresponds to the involution on exchanging the two factors. This involution leaves the Kirwan map, and of the previous section invariant. This implies that all of these functions descend to functions on which also parametrizes the -orbits on .
4.2. The group case
As a first example, we consider equipped with the action of by left and right multiplication, as described in Example 1.1(5). In particular, here we identify with the homogeneous space . We choose the Borel subgroup where is a Borel subgroup of and is its opposite. By the Peter-Weyl Theorem, decomposes as a -module as follows:
where is the set of dominant weights of . It follows from Definition 1.7 that and that the lattice of as a -spherical variety is the sublattice . Using Theorem 1.12, [Timashev, Theorem 24.2], [Timashev, Lemma 24.3] and the Cartan decomposition, it can then be shown that the valuation cone can be identified with the antidominant Weyl chamber in , i.e., .
In the case of , to which we now restrict, we can also give an explicit description of the spherical functions. We choose the basis for where denotes the usual -th fundamental weight of . The corresponding irreducible -representation is . To construct the spherical function corresponding to , we need to find an isomorphic copy of in , considered as an -representation. Let be subsets of of cardinality . Let denote the minor of an matrix corresponding to the subsets and , i.e., the determinant of the submatrix with rows in and columns in . We view as an element in . By explicitly analyzing the action of the Chevalley generators (for positive roots ) as well as by computing the -weights of the , it is not difficult to see that the span of the is a subrepresentation of isomorphic to . Moreover, it is straightforward to see that each is a -weight vector, and that there exists a -invariant inner product with respect to which the form an orthonormal basis. Thus we have
where the normalization factor ensures that .
Recall the singular value decomposition theorem, which states that any can be expressed as a product
where and is a diagonal matrix with positive real entries whose product is equal to . Let and consider the spherical logarithm map on corresponding to . Since the spherical function is -invariant by construction, it follows that
(Note that the components of the function are exactly the symmetric functions on the eigenvalues of , i.e., the coefficients of the characteristic polynomial of .)
For the special case , an elementary calculus (maximization) computation shows that the resulting region is bounded by
from which it can be seen that as approaches , the spherical amoebae approach . See Figure 1. Also compare with [Kaveh-Manon, Example 7.7].

4.3. The basic affine space
Consider the spherical homogeneous space where is a maximal unipotent subgroup (see Example 1.1(4)). One knows that is a quasi-affine variety. The ring of regular functions can be identified with the ring of (right) -invariants. Thus, as a -module it decomposes as:
Moreover, the multiplication in the algebra corresponds to Cartan multiplication between the . It follows that the tail cone consists of the origin only and hence the valuation cone is the whole vectors space .
Let us give an explicit description of the spherical functions on in the case of . One can take to be the subgroup of upper triangular matrices with ’s on the diagonal. Set and consider the natural action of on the exterior algebra . If are the standard basis vectors of , then the stabilizer of is . As is a locally closed embedding with quasi-affine image, we obtain a surjective map and it straightforwardly follows that the algebra is generated by the so-called flag minors: Let . For subsets with , let denote the minor of which is the determinant of the submatrix of with rows corresponding to and columns corresponding to . A flag minor is a minor of the form , that is when corresponds to the first columns of .
Let . There are spherical functions corresponding to the fundamental weights . The representation corresponding to the -th fundamental weight is and the flag minors are a vector space basis consisting of weight vectors. Consider the standard Hermitian product on . This induces Hermitian products on for al . One verifies that, for each , the vector is an orthonormal basis for . It follows that the flag minors are an orthonormal basis for . Thus the -th spherical function corresponding to is given by:
where the sum is over all subsets with . Note that . Consider the spherical logarithm map corresponding to our choice of . We have
with , and thus the image of is the whole space and agrees (for all ) with the valuation cone.
4.4. Example of
Throughout this section, we let and for . The variety is a very well-known example of a spherical homogeneous space. Nevertheless, for the convenience of the reader we try to give details of the proofs in this section.
The maximal compact subgroup of is which we realize explicitly as
where denotes the conjugate transpose. Let act on as follows:
(15) |
We denote the standard basis of by and the coordinates on the two factors by and respectively. With respect to the action (15) it is straightforward to compute that the stabilizer of is
Lemma 4.1.
where we use the notation and .
Proof.
The inclusion “” is straightforward, since for any , it immediately follows from the definition of the action that would have the property .
To see the inclusion “”, suppose that satisfies . We wish to find an element in that takes to . First notice that since , there must exist some matrix with the property that the top row of is . Now define a matrix by replacing the first column of with (a column vector). We claim that is an element in and that . By construction we have , so it remains to see that and that .
We claim . To see that we make an explicit computation. Given an matrix we denote by the matrix obtained by deleting its -th row and -th column. Below we compute by expansion along its left-most column, which by construction is .
We next claim that , or in other words, that has top row equal to . By a Cramer’s rule argument similar to that above, we can see that the -th matrix entry of is the -th cofactor of , which is equal to the -th cofactor of , which in turn is equal to the -th entry of . But this last value is exactly by construction of , so we are done. ∎
It follows from the above lemma that the homogeneous space is an affine variety and its coordinate ring may be identified as
Let be the Borel subgroup of upper triangular matrices in and the maximal torus consisting of all diagonal matrices in . Let be the maximal unipotent subgroup of consisting of upper triangular matrices with ’s on the diagonal. We show next that the homogeneous space is in fact spherical, i.e., has a dense open -orbit. We have the following.
Lemma 4.2.
The subset
is a dense open -orbit in . In particular, is a spherical homogeneous space.
Proof.
The inclusion “” is straightforward. To show the reverse inclusion, let with and . Consider the matrix
We claim that is in the Borel subgroup of and that . By construction is upper-triangular, and from the diagonal entries it easily follows that , so is in . It is straightforward to see from the definition of that . To see that , it suffices to see that the top row of is . Moreover, since , the last entry is determined by and , it in fact suffices to check the first entries. Fix with . We wish to show that the -th entry of is equal to . It is straightforward to check this explicitly using the adjoint form of as in the proof of the previous lemma and the explicit formula for given above. ∎
We now describe parts of the Luna-Vust data associated to the spherical homogeneous space . Recall that the colors of is defined to be the set of -invariant prime divisors of . From Lemma 4.2 we know the open -orbit is defined by the condition , so any -invariant prime divisor is contained in its complement .
Lemma 4.3.
The two elements are prime elements in .
Proof.
To see that is a prime element, it would suffice to show that is a domain. Note that there is an isomorphism
so it suffices to show that the RHS is a domain. Since the polynomial is of degree , if it factors non-trivially then it must be of the form where so both and are linear polynomials in (with constant term). It is not hard to see by direct computation that this is impossible, so is irreducible and hence the RHS of the equality above is an integral domain. We conclude is prime in , as desired. The argument for is similar. ∎
From Lemma 4.3 it follows that and are prime divisors. Moreover, it is immediate from (15) that both are -invariant. It follows that are the two colors in , i.e., .
Remark 4.4.
Note that is not contained in the open -orbit and the colors consist of several -orbits. Indeed, we have the following “interesting” -orbits:
-
•
is a base point of the open -orbit in .
-
•
is the open -orbit of the color .
-
•
is the open -orbit of the color .
Note that the colors are each a union of multiple -orbits.
Next, we need to compute the -semi-invariants in the ring . We need some preparation.
Lemma 4.5.
The coordinate ring is a UFD, a unique factorization domain. Moreover, the units in consists of the non-zero constants, i.e. .
Proof.
Consider the localization of by the prime element and observe that we have an isomorphism
since . In particular, the ring on the right hand side is a UFD. From [Eisenbud, Lemma 19.20] we know that if is an integral domain and a prime element and the localization is a UFD, then is a UFD. Applying this to and , we conclude that is also a UFD.
Now we wish to prove the claim about the units. Embed into the localization . In particular, any unit in must also be a unit in the localization. The units of are of the form for and . However, is a prime element in , so it is not a unit in . Thus the only units in are the non-zero constants. ∎
Recall that a function is called -semi-invariant if there exists a character such that for all . The set of -semi-invariants is denoted . From the above lemma we can deduce the following.
Corollary 4.6.
We have
Proof.
Let be non-zero and -semi-invariant. Then the vanishing locus of must be contained in since cannot vanish on the dense open -orbit (see Lemma 4.2). Thus for some non-negative integers and and is an invertible regular function on . In particular, it is a unit in . We saw in Lemma 4.5 that the units are the non-zero constants, so we conclude that for some , as desired. ∎
We can now compute . Let be the fundamental weights associated to our choice , i.e., for we have
It is straightforward to compute that the -weights of and are and respectively. The association gives a map whose image is denoted . If then so is a semigroup. In our case , the above computation shows that is freely generated by and , i.e. . For what follows, we choose , with associated -semi-invariant functions . Note that is also a basis of , and with respect to this choice of basis, the map of (2) becomes
We have the following, where we temporarily work over , the most natural setting for this lemma.
Lemma 4.7.
The image of under is
(16) |
Proof.
We first show that the LHS is contained in the RHS. Let . For any with , there is a permutation of the coordinates such that . A similar statement holds for and . Since is -invariant, it follows that
On the other hand, since in , we obtain from the axioms of valuations that
Hence lies in the RHS of (16), as desired.
We now claim that the RHS is contained in the LHS. By Theorem 1.5 the image is a (rational) polyhedral cone, in order to show that the RHS is contained in the LHS, it suffices to show that integral points in the RHS are contained in the LHS. Let with . Consider the subvariety
and consider the -action on given as the product of the given -action on and the trivial action on the last factor. This induces an -action on : indeed, if , then for , we have . Note that is -invariant, so is also invariant. There is an -equivariant projection
which induces an inclusion of the corresponding function fields . We claim that is a prime divisor in . Indeed, is a prime element, since
and the ring on the RHS can be seen to be an integral domain from the fact that is irreducible in by an argument similar to the proof of Lemma 4.3. Thus we may define to be the geometric valuation on given by the order of vanishing of a rational function along the (prime) divisor . Since is -invariant, this is an -invariant valuation, and the restriction to (the image of) yields an -invariant valuation in . We compute
Thus we have shown that lies in the image under of , as desired. ∎
As discussed above, in order to compare the valuation cone with the image of the spherical logarithm we should tensor with and consider . It is clear that the closure is and this is what we consider as the valuation cone in Proposition 4.9 below.
The above discussion can be summarized in the following picture. The gray half-space defined by is the valuation cone . We have also indicated the images under of the (geometric valuations corresponding to the) two colors and ; note that these do not lie in the valuation cone since they are not -invariant (only -invariant).
Next we compute the spherical functions. Consider the decomposition . Let the natural representation of and its dual representation. Then has highest weight while the dual representation has highest weight . The standard scalar product turns the standard bases into unitary bases on and respectively. Note that and are naturally embedded in by the identifications and . The variables and form weight bases (with respect to ) of and respectively. Hence, we can define, following (6) in Section 2, the following spherical functions:
(17) |
both of which lie in .
Remark 4.8.
Note that and are both -invariant, because and are orthonormal and the action of leaves lengths invariant.
Finally, we analyze the spherical amoebas, i.e. the images of under the spherical logarithm maps, and explicitly compute their limit as . In our case of we can prove that the limit is precisely equal to the valuation cone, see Proposition 4.9. Recall that our spherical logarithm map is given by
From (17) we can also express this more explicitly as
In this special case, an explicit computation yields the following.
Proposition 4.9.
Let and . Then for all . In particular, as approaches , the sets converge to the valuation cone in the sense of Kuratowski.
Proof.
Let , i.e., . Let with . Then which implies that is a strictly decreasing function on . Using this, we can make the following computation:
where the first inequality uses the standard Cauchy-Schwartz inequality together with the fact that is a decreasing function. It follows that . It remains to show that . Let , i.e., with . Then , and, as is a strictly decreasing function for , we get that or equivalently . Hence is defined. We set and and observe that . Finally, we may compute that
so