Startseite Combinatorics of stratified hyperbolic slices
Artikel Open Access

Combinatorics of stratified hyperbolic slices

  • Arne Lien
Veröffentlicht/Copyright: 19. Juli 2025
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

We study sets of univariate hyperbolic polynomials that share the same first few coefficients and show that they have a natural combinatorial description akin to that of polytopes. We define a stratification of these sets in terms of root arrangements of hyperbolic polynomials and show that any stratum is either empty, a point or of maximal dimension, and in the latter case we characterise its relative interior. This is used to show that the poset of strata is a graded, atomic and coatomic lattice and to provide an algorithm for computing which root arrangements are realised in these sets of hyperbolic polynomials.

MSC 2010: 26C05; 12D10

The topic of this article is the study of (canonical) hyperbolic slices which are sets of univariate hyperbolic polynomials that share the same first few coefficients. The motivation to study such sets stems from the article [14], where they were used to provide a new proof of Timofte’s degree and half degree principle for symmetric polynomials. Therefore these sets are deeply connected with the study of multivariate symmetric polynomials.

A natural way of describing the root arrangement of a hyperbolic polynomial is to construct its partition of multiplicities. However, by also considering in which order the roots arise we get a finer description of the polynomial’s root arrangement which we call its composition. We shall see that by using compositions to stratify the set of hyperbolic polynomials, we get a lattice of strata that is graded, atomic and coatomic.

To establish these combinatorial properties of the set of strata we show that the strata are connected to a type of symmetric, real algebraic sets called Vandermonde varieties. This connection allows us to show that the relative interior of a stratum consists of the polynomials with the largest number of distinct roots and that a stratum is either empty, a point or of the generic dimension of a nonempty Vandermonde variety.

We begin in Section 1 by introducing the sets that we study and by defining our stratification. We look into an example of a set of degree 5 hyperbolic polynomials with the same first three coefficients (Figure 1), and we show in Proposition 1.6 that the collection of strata, partially ordered by inclusion, is a lattice.

Figure 1 
				Hyperbolic slice H2(f) for Example 1.4 below
Figure 1

Hyperbolic slice H2(f) for Example 1.4 below

In Section 2 we show that a result from [10] on Vandermonde varieties can be used to describe the relative interior of the strata (Theorem 2.6) and to show that the strata are either empty, a point or of maximal dimension (Theorem 3.10, which says that the poset of strata is a graded, atomic and coatomic lattice. Finally this leads us to Algorithm 3.12 that determines which compositions occur in our sets.

We finish with Section 4 where we discuss how a result from [12] on discriminants and subdiscriminants implies that the boundaries of the sets of hyperbolic polynomials have a concave-like property as exhibited in the picture above. In combination with Theorem 3.10, this leads to the open question of whether hyperbolic slices are concave variations of polytopes.

1 Stratification

We start off by defining hyperbolic slices and showing how we stratify these sets. Then we look closer at the example from the introduction. We finish this section by showing that the collection of strata, partially ordered by inclusion, is a lattice.

Definition 1.1

A univariate polynomial f ∈ ℝ[t] is called hyperbolic if all its roots are real.

Let d be some positive integer throughout this article and let H denote the subset of ℝ[t] consisting of the hyperbolic polynomials of degree d. For fH we denote by Hs(f) the set of all hyperbolic polynomials of degree d with the same s + 1 first coefficients as f. That is, if f=f0td+f1td1++fdH, then

H s ( f ) : = h 0 t d + h 1 t d 1 + + h d H h i = f i  for all  i s .

If a = (a1, . . . , ad) are the roots of f, then it is well known that fi = (−1)iei(a) where ei is the ith elementary symmetric polynomial in d variables. Thus the subscript s refers to the number of fixed elementary symmetric polynomials.

Let f be some monic hyperbolic polynomial of degree d ≥ 1 throughout this article. We refer to Hs(f) as a (canonical) hyperbolic slice, where we omit the word canonical as these are the only type of hyperbolic slices we consider. The general definition can be found in [15].

Often a polynomial h=td+h1td1++hdHs(f) will be identified with the point hs+1,,hdds without specifying this change of basis. Thus when we consider topological questions, we equip Hs(f) with the subspace topology of the Euclidean topology on ℝds.

To stratify the hyperbolic slices we introduce compositions and a corresponding partial order.

Definition 1.2

A composition of d is a tuple of positive integers, u = (u1, . . . , ul), with

i = 1 l u i = d .

The integers ui are called the parts of u, and ℓ(u) := l is the length of u.

We often use the shorthand (1d) for the composition whose parts are all equal to 1. Also, when nothing further is specified, we let u denote a composition of d.

Definition 1.3

Let u and v be two compositions of a positive integer d. Then v < u if v can be obtained from u by replacing some of the commas in u with plus signs.

To any hyperbolic polynomial f of degree d, we associate a composition of d in the following way: let a1 < a2 ⋅⋅ ⋅ < al be the distinct roots of f with respective multiplicities m1, . . . , ml. Then the composition of f is the tuple v(f) := (m1, . . . , ml), which is a composition of d.

The strata we will look at are given by

H s u ( f ) : = h H s ( f ) v ( h ) u .

Note that Hs1d(f)=Hs(f) since any composition of d distinct from (1d) is smaller than (1d).

Example 1.4

Let d = 5 and s = 2 and let

f = ( t + π ) ( t + 2 ) t ( t 1 , 23456789123456789 ) ( t e ) .

If we map the last three coefficients of the polynomials in H2(f), we get the three-dimensional Figure 1 in the introduction. The polynomials with no repeated roots make up the interior, and the strata H2u(f) with u ≠ (1d) lie on the pieces of the boundary as indicated by Figure 2. For instance, if u = (2, 1, 2), then the compositions (2, 3) and (3, 2) are the compositions smaller than u; thus H2u(f) corresponds to a closed one-dimensional piece of the boundary (including its endpoints).

Figure 2 
							Strata H2u(f)
										$H_{2}^{u}(f)$
									in H2(f)
Figure 2

Strata H2u(f) in H2(f)

Next we show that the poset of strata is a lattice. If a and b are elements of a lattice, we denote their join by ab and their meet by ab.

Lemma 1.5

The poset of compositions of d is a lattice.

Proof. We prove this by explicitly constructing the join of two compositions u and v of d. Having done so, the existence and uniqueness of the meet is also established since the meet of u and v is just the join of all compositions smaller than both u and v.

If a and b are two compositions of d, note that ab if and only if

a 1 , a 1 + a 2 , , d b 1 , b 1 + b 2 , , d .

So let M = {u1, u1 + u2, . . . , d} ∪ {v1, v1 + v2, . . . , d} and construct the tuple m = (m1, m2, . . . , ml) containing all the distinct elements of M where mi < mi+1 for any i ∈ [l − 1]; note that ml = d.

Consider the composition w = (m1, m2m1, m3m2, . . . , mlml−1) of ml = d. Since {u1, u1+u2, . . . , d} ⊆ M and {v1, v1+v2, . . . , d} ⊆ M, we have uw and vw. By construction, M is the unique minimal set that contains both {u1, u1 + u2, . . . , d} and {v1, v1 + v2, . . . , d}, hence w is the join of u and v.

From Lemma 1.5 we immediately get that the strata of Hs(f), partially ordered by inclusion, form a lattice. To see this we determine the meet of two strata Hsu(f) and Hsv(f) of Hs(f). The meet of two strata is contained in their intersection since the partial order is given by inclusion. Also, by definition of the strata we have

H s u ( f ) H s v ( f ) = H s u v ( f ) ,

so the intersection is a stratum of Hs(f). Thus we have that

H s u ( f ) H s v ( f ) = H s u v ( f )

and we have shown the following:

Proposition 1.6

The set of strata of Hs(f), partially ordered by inclusion, is a lattice.

However, it is worth pointing out that just because the meet of Hsu(f) and Hsv(f) is Hsuv(f), this does not a priori mean that uv is the only composition w such that Hsu(f)Hsv(f)=Hsw(f). For instance, if Hsu(f) is empty then Hsu(f)Hsv(f)=Hsu(f) even if uvu. So different compositions may label the same stratum and this does indeed happen sometimes.

A similar problem can arise when we consider joins of strata. However in this case, we shall see in Section 3 that this can be overcome by requiring u and v to be the minimal compositions that can label the strata Hsu(f) and Hsv(f). Let us look at an example of these oddities:

Example 1.7

We return to the tetrahedron-like Example 1.4. Let hH2(f) be a polynomial with composition (2, 3). Then we can see that the compositions occurring in H3(h) are

1 5 , ( 1 , 1 , 1 , 2 ) , ( 1 , 1 , 2 , 1 ) , ( 1 , 2 , 1 , 1 ) , ( 1 , 2 , 2 ) , ( 1 , 3 , 1 )  and  ( 2 , 3 ) .

So the strata labeled by the compositions (1, 1, 3), (2, 1, 2), (2, 2, 1) and (2, 3) are all the same stratum.

For a converse example, consider the stratum labelled by (1, 2, 2). If we look at the join of H3(1,2,2)(h) and H3(2,2,1)(h)=H3(2,3)(h), then this is equal to H3(1,1,2,1)(h). But the join of (1, 2, 2) and (2, 2, 1) equals (15) and the corresponding stratum is H315(h)=H3(h), which is not the join of H3(1,4)(h) and H3(2,2,1)(h).

However, the minimal composition that can label the stratum H3(2,2,1)(h) is (2, 3). The join of (1, 2, 2) and (2, 3) equals (1, 1, 2, 1), so by choosing to label H3(2,2,1)(h) by the minimal composition that may label the stratum we avoid this particular peculiarity.

2 Vandermonde varieties

To describe the combinatorial structure of the lattice of strata we need to establish some geometric properties of our strata. In particular, we will show that Arnold’s, Givental’s and Kostov’s work on so-called "Vandermonde varieties" (see [1], [9] and [10]) implies that Hsu(f) is either empty, or a point or of dimension ℓ(u) − s, and that in the latter case the polynomials with composition u make up the relative interior.

To see the connection to their work, let the symmetric group Sym([d]) act on ℝ[x1, . . . , xd] by permuting the variables. It is well known that the elementary symmetric polynomials generate the ring of invariants; see Chapter 7.1 in [6]. The induced action on ℝd permutes the coordinates of the points in ℝd, and the orbit space ℝd/Sym([d]) can be identified with the image of ℝd under the mapping Π : ℝdH, where

Π ( x ) = t d e 1 ( x ) t d 1 + + ( 1 ) d e d ( x )

and ei(x) is the ith elementary symmetric polynomial.

Thus the orbit space can be identified with the monic hyperbolic polynomials of degree d, and by restricting Π to the set Kd := {x ∈ ℝd | x1 ≤⋅ ⋅⋅ ≤ xd} we obtain a bijection between Kd and Π(ℝd). So we see that hyperbolic slices can be thought of as sections of the orbit space obtained by intersecting it with certain affine hyperplanes.

Similarly, if u = (u1, . . . , ul) and hHsu(f) has the roots a1 ≤⋅ ⋅⋅ ≤ al , then

h ( t ) = t d e 1 a u t d 1 + e 2 a u t d 2 + + ( 1 ) d e d a u ,

where au := (a1, . . . , a1, . . . , al , . . . , al) and ai is repeated ui times. So if we define the real algebraic set

V s u ( f ) : = x l e i x u = ( 1 ) i f i  for every  i [ s ] ,

then Hsu(f) is the image of Vsu(f)Kl under the mapping

Π u ( x ) : = t d e 1 x u t d 1 + + ( 1 ) d e d x u .

Another set of generators for the invariant ring ℝ[x1, . . . , xd]Sym([d]) is the set of all power sums pi(x) := j[d]xji with i ∈ [d]. Newton’s identities express the first s power sums as polynomials in the first s elementary symmetric polynomials, hence we can equivalently define Vsu(f) as {x ∈ ℝl | pi(xu) = ci for every i ∈ [s]}, where c1, . . . , cs ∈ ℝ are obtained from −f1, . . . , (−1)s fs using Newton’s identities. Such sets are referred to as Vandermonde varieties, since the Jacobian of the first s power sums is a constant multiple of a Vandermonde determinant.

To make use of the previous work on Vandermonde varieties, we first need to establish that Πu is a homeomorphism. Note that, as with Hs(f), we equip the sets Vsu(f),Kl and Hsu(f) with the subspace topology of the Euclidean topology.

Let Bε(a) denote the real open ball about a ∈ ℝk with radius ε > 0, and let Bε(a)¯ denote its closure. Similarly, let Dε(z) denote the complex open ball about z ∈ ℂk with radius ε, and let Dε(Z)¯ denote its closure.

Lemma 2.1

Let ℓ(u) = l, then Hsu(f) is closed inds and

Π u : V s u ( f ) K l H s u ( f )

is a homeomorphism.

Proof. Note that Πu is a bijection and a polynomial mapping, thus it is a continuous bijection. To see that the inverse map is continuous and that Hsu(f) is closed in ℝds, we show that the images of closed sets in Vsu(f)Kl are closed in ℝds.

Let S be a closed subset of Vsu(f)Kl; then since Vsu(f) and Kl are closed in ℝl , so is S. Let

ι u : V s u ( f ) K l d

be the inclusion x1,,xlxu=x1,,x1,,xl,,xl, where xi is repeated ui times. Then Πu(x) = (Πιu)(x) and clearly ιu(S) is a closed subset of ℂd.

Let h = td +h1td−1 +⋅ ⋅ ⋅+hdΠu(S) have the roots a = (a1, . . . , ad) and let hi = fi for all i ∈ [s]. Let ε > 0 be such that Dε(σ(a))∩ ιu(S) is empty for every σ ∈ Sym([d]). If b1, . . . , bk are the distinct roots of h with respective multiplicities v1, . . . , vk, then there is a δ > 0 such that any polynomial g of degree d with |higi | ≤ δ for all i ∈ [d] has exactly vi zeroes in Dε(bi); this statement is proved in [18]. Thus, since Dε(σ(a)) ∩ ιu(S) is empty for any σ ∈ Sym([d]), then so is Bδ(h) ∩ Πu(S), and therefore Πu(S) is closed in ℝds. □

Proposition 2.2

The sets Vsu(f)Kl and Hsu(f) are contractible or empty.

Proof. Firstly, we can use Newton’s identities to define Vsu(f) in terms of the first s power sums in d variables. Then the proof that Vsu(f)Kl is contractible or empty can be found in [10, Theorem 1.1]. By Lemma 2.1 the map Πu:Vsu(f)KlHsu(f) is a homeomorphism, thus Hsu(f) is contractible if it is nonempty. □

To see how this proposition can be used to further describe our strata we need some more definitions. First note that as Hsu(f) is the image of a semialgebraic set under a polynomial mapping, it is semialgebraic. Thus the dimension of the stratum is the maximum integer n such that Hsu(f) contains an open set which is homeomorphic to an open set of ℝn; see [4, Chapter 2.8].

Definition 2.3

If Hsu(f) is a nonempty stratum and n=dimHsu(f), then

  • the relative interior RelInt Hsu(f) of Hsu(f) is the set of all polynomials hHsu(f) such that some open neighbourhood of h is homeomorphic to an open set in ℝn, and

  • the relative boundary RelBdHsu(f) of Hsu(f) is the set of all polynomials in Hsu(f) which are not in the relative interior.

We can use Proposition 2.2 to describe the relative interior and the relative boundary of our strata and also to determine their dimensions. But the first consequence of the proposition that we need is the following:

Lemma 2.4

If ℓ(u) ≤ s, then Hsu(f) contains at most one polynomial.

Proof. Suppose that hHsu(f) has the distinct roots a = (a1, . . . , ak) and composition v = (v1, . . . , vk); then k ≤ ℓ(u) ≤ s. If k = 1, then v = (d) and there is only one solution to the equation

e 1 x v = d x 1 = f 1 ,

and so we have Hsv(f)=Hkv(f)={h}. If k > 1, then as previously mentioned, Vsv(f) can be defined as

x k p 1 x u = c 1 , , p k x u = c k ,

where pi is the ith power sum in d variables and c1, . . . , ck ∈ ℝ are obtained from the numbers −f1, . . . , (−1)k fk using Newton’s identities.

The map F : ℝk → ℝk where F(x) = (p1(xv), . . . , pk(xv)) is a continuously differentiable function whose Jacobian matrix is ivjxji1i,jk and so its determinant is

i = 1 k i v i 1 j < r k x j x r .

Since all the ai’s are distinct, the determinant is nonzero at a. Thus the Jacobian matrix is invertible and by the inverse function theorem, F is invertible on some neighbourhood U of F(a) = (c1, . . . , ck). By Proposition 2.2, Vkv(f)Kk is contractible and since a is isolated in this set, it must be the only point there. Therefore we have Hsv(f)=Hkv(f)={h}.

So for any composition wu that occurs in Hsu(f) we have that Hsu(f) is a point. Since there are only finitely many compositions smaller than or equal to u,Hsu(f) contains only finitely many points. But since Hsu(f) is contractible it can contain at most one point. □

Let

P r : k k r

denote the projection that forgets the last r coordinates. This map will help us describe the relative interior.

Proposition 2.5

If l = ℓ(u) > s, then the map

P d l : H s u ( f ) l s

is a homeomorphism onto its image and the image is closed inls.

Proof. Firstly we consider the case when l = 1; then u = (d) and s = 0. So for any a ∈ ℝ we have that (ta)d=tddatd1++(a)dH0u(f). Thus Pd−1((ta)d) = tddatd−1 and so the map

P d l ° Π u :

is essentially just mapping a to −da. This is naturally a homeomorphism and since, by Lemma 2.1, Πu is a homeomorphism, then so is Pdl. Lastly, since the image of Pdl is all of ℝ, the image is closed in ℝ.

Next suppose that l ≥ 2. By Lemma 2.4, the polynomials of Hsu(h) are uniquely determined by their first l coefficients, thus Pdl is a bijection between Hsu(h) and PdlHsu(f). Also, the topology on Hsu(f) is the subspace topology of the product topology on ℝds with respect to the projections on each coordinate. Thus, by definition, the map Pdl is continuous.

To see that the inverse is continuous and that PdlHsu(f) is closed, we will show that the images of closed subsets of Hsu(f) are closed in ℝls. So let S be a closed subset of Hsu(f); then S is also a closed subset of ℝds since, by Lemma 2.1, Hsu(f) is closed in ℝds.

Let g be a point in the closure of Pdl(S). Then for any ε > 0, the closed ball B¯ ε(g) meets Pdl(S) and so the inverse image

P d l 1 B ¯ ε ( g ) P d l ( S ) = B ¯ ε ( g ) × d l S

is nonempty. It is also closed since B¯ ε(g)×dl and S are closed in ℝds.

Since Πu is a homeomorphism, the set

M = Π u 1 B ¯ ε ( g ) × d l S

is a closed subset of Vsu(f)Kl, and since Vsu(f)Kl is closed in ℝl, so is M. We also have that

M Π u 1 ( S ) x K l g i ε ( 1 ) i e i x u g i + ε  for all  i [ l ] .

So if aM, then e1(au)≤ g1ε and e2(au)≥ g2ε since l ≥ 2. Thus, by Newton’s identities, we have

p 2 a u = e 1 2 a u 2 e 2 a u g 1 2 2 g 1 ε + ε 2 2 g 2 + 2 ε

and so M is bounded. Since M is closed and bounded, it is compact.

Since Πu and Pdl are continuous, Pdl°Πu is continuous. And since the continuous image of a compact set is compact, we have that Pdl°Πu(M) is compact. Thus

g P d l Π u ( M ) P d l ( S )

and so Pdl(S) is closed in ℝls. Therefore Pdl is a closed map and thus Pdl is a homeomorphism. Lastly, by setting S=Hsu(f), we see that PdlHsu(f) is closed in ℝls.

We see from Lemma 2.5 that when ℓ(u) ≥ s, the largest dimension that Hsu(f) can have is ℓ(u) − s. Thus we say that the maximal dimension of Hsu(f) is max{ℓ(u) − s, 0}.

Theorem 2.6

If Hsu(f) contains a polynomial with composition u, then Hsu(f) is maximal dimensional and its relative interior consists of the polynomials with composition u.

Proof. If s = 0 and l = ℓ(u), the map PdlΠu:KlPdlHu(f) is a homeomorphism by Lemma 2.1 and Lemma 2.5. So since the dimension of K l is l and its interior points are the points with no repeated coordinates, the dimension of PdlH0u(f)l is l and its interior points are the polynomials with composition u.

Next suppose that s > 0 and let Afir denote the affine hyperplane of ℝr defined by fixing the ith coordinate to be equal to fi. Then

P d l H s u ( f ) = P d l H 0 u ( f ) A f 1 d A f s d = P d l H 0 u ( f ) A f 1 l A f s l .

Thus if there is a polynomial hHsu(f) with composition u, then Pdl(h) lies in the interior of PdlH0u(f) and therefore also in the interior of PdlH0u(f) and PdlH0u(f) must be of dimension max{ls, 0}. Since Pls is a homeomorphism, Hsu(f) is maximal dimensional and h is in its relative interior.

For the reverse inclusion, suppose that l > s so that Hsu(f) is at least one-dimensional. If its relative interior contains a polynomial g with v(g) < u, then Pdl(g) lies in the interior of the (ls)-dimensional set PdlHsu(f) ls. Thus Pdl(g) lies in the interior of the one-dimensional set PdlHl1u(g).

By Lemma 2.4 there are only finitely many polynomials in Hl1u(g) with a smaller composition than u. Thus there are two polynomials p and p+ in Hl1u(g) with composition u, and a δ > 0 such that

P d l p = P d l ( g ) δ  and  P d l p + = P d l ( g ) + δ .

Since Pdl(p) and Pdl(p+) are interior points of PdlH0u(f), there is an ε > 0 such that BεPdlp and Bε(Pdl(p+)) are contained in the interior of PdlH0u(f). And since Pdl(g) is in the boundary of PdlH0u(f), the ball Bε(Pdl(g)) must contain a polynomial q = td + q1td−1 + ⋅⋅ ⋅ + ql tdl that is not in PdlH0u(f).

Thus Aq1lAql1l is a line that passes through q and the two balls Bε(Pdl(p)) and Bε(Pdl(p+)). But if q separates the nonempty sets

B ε P d l p A q 1 l A q l 1 l

and

B ε P d l p + A q 1 l A q l 1 l ,

then

P d l H 0 u ( f ) A q 1 l A q l 1 l = P d l H l q 1 u p +

is nonempty but not contractible. This contradicts Proposition 2.2; therefore g cannot be in the relative interior of Hsu(f).

Thus if ℓ(u) > s, then the stratum Hsu(f) is of maximal dimension if and only if it contains a polynomial with composition u. We can use this observation to determine all the possibilities for the dimension of Hsu(f). It is worth mentioning that a similar observation can be found in [3], Proposition 5.

Theorem 2.7

If ℓ(u) > s and Hsu(f) contains a polynomial with at least s distinct roots, then Hsu(f) is maximal dimensional. If not, then Hsu(f) is either empty or a single polynomial.

Proof. If s = 0, then any composition occurs and thus by Theorem 2.3, any stratum is maximal dimensional. Similarly, if s = 1, then for any composition u, the polynomial −e1(xu)− f1 has a real zero with l = ℓ(u) distinct coordinates ordered increasingly. To see this pick l real numbers a1, . . . , al such that a1/u1 < ⋅ ⋅⋅ < al/ul and

let

a = f 1 i a i a 1 u 1 , , a l u l ;

then

e 1 a u = j u j f 1 i a i a j u j = f 1 i a i j a j = f 1 .

Thus the composition u occurs and Hsu(f) is of maximal dimension.

Next we suppose that s ≥ 2. If ℓ(u) ≤ s or Hsu(f) does not contain a polynomial with at least s distinct roots, then Hsu(f)= wu and (w)=s1Hsw(f). By Lemma 2.4, Hsw(f) contains at most one point when ℓ(w) = s − 1. Since there are only finitely many compositions of length s − 1, the stratum Hsu(f) contains only finitely many polynomials and since Hsu(f) is contractible it contains at most one polynomial.

So suppose that hHsu(f) has ks distinct roots and v(h) < u. Let vu be a composition that covers v(h). Then ℓ(v) = k + 1 and so by Proposition 2.5, Hkv(h) is at most one-dimensional. Since v(h) < v we can write h as i=1k+1taivi and without loss of generality we may assume that a1 < ⋅ ⋅⋅ < ak = ak+1.

As in the proof of Lemma 2.4, if we define Vkv(f) as xk+1p1xu=c1,,pkxu=ck, then the Jacobian matrix of the defining polynomials is ivjxji1ik,jk+1 and the determinant of the leftmost k × k submatrix is

i = 1 k i v i 1 j < r k x j x r .

Since the first k coordinates of a = (a1, . . . , ak+1) are distinct, the determinant does not vanish at aVkv(h). So by Proposition 3.3.10 in [4], a is a nonsingular point of a one-dimensional irreducible component, say V, of Vkv(h). Thus a lies in an open neighbourhood U of V where U is a one-dimensional manifold.

By Lemma 2.4, the one-dimensional manifold U intersects the hyperplane H = {x ∈ ℝk+1 | xk = xk+1} only once. So U must meet the open halfspace H+ := {x ∈ ℝk+1 | xk < xk+1} and thus there is a point in Vkv(h) with no repeated coordinates. So Hkv(h)Hsu(f) contains a polynomial g with composition v.

We can apply the same argument to the polynomial g if v < u, and keep doing this inductively until we find a polynomial with composition u. Then by Theorem 2.6, Hsu(f) is maximal dimensional. □

Corollary 2.8

Any stratum equals the closure of its relative interior.

Proof. By Theorem 2.7 we may suppose that Hsu(f) is maximal dimensional and at least one-dimensional. We will prove the statement by induction on the dimension of the strata. If Hsu(f) is one-dimensional, then it is connected by Proposition 2.2. Thus it contains at most two relative boundary points, and any open ball about a point of Hsu(f) must contain infinitely many points from the relative interior. Thus Hsu(f) is the closure of its relative interior.

Now suppose that the statement is true for all (n − 1)-dimensional strata, where n − 1 ≥ 1. Suppose that Hsu(f) is n-dimensional and hHsu(f). By Proposition 2.2, Hsu(f) is connected, so for any ε > 0 the ball Bε(h) contains infinitely many polynomials from Hsu(f). Since there are only finitely many polynomials in Hsu(f) with at most s distinct roots, there is a gBε(h)Hsu(f) with at least s + 1 distinct roots.

Then by Theorem 2.7 the stratum Hs+1u(g) is (n − 1)-dimensional, and by the inductive hypothesis g is in the closure of its relative interior. So by Theorem 2.6, any open ball about g contains a polynomial with composition u. Thus Bε(h) also contains a polynomial with composition u, and so h is in the closure of the relative interior of Hsu(f).

3 Combinatorial structure

In this section we explore the combinatorial structure of the lattice of strata and show that this lattice is graded, atomic and coatomic. Then we use this to determine which compositions actually occur in a given hyperbolic slice. Due to Theorem 2.7 we may assume that H s(f) is of dimension ds > 0 for this section since the one-point lattice trivially satisfies the main combinatorial properties that follow. We begin with the notion of a graded poset.

Definition 3.1

A totally ordered subset of a poset is a chain, and if a chain is maximal with respect to inclusion it is a maximal chain. A poset in which every maximal chain has the same length is called graded.

To see why we call such a poset graded, let y0 < ⋅ ⋅⋅ < yl and z0 < ⋅ ⋅⋅ < zl be two maximal chains of a graded poset L where yi = zj, for some i and j. Then we have i = j, otherwise y0 < y1 < ⋅ ⋅⋅ < yi = zj < zj+1 < ⋅ ⋅⋅ < zl is a maximal chain which is not of length l +1 contradicting the gradedness of L. Thus the rank of yi, rank(yi) := i, is well defined and we can write the poset as the disjoint union L=k0L(k), where the elements of L(k) are the rank k elements.

Lemma 3.2

If s ≥ 2 and Hsu(f) is at least one-dimensional, then it is compact and its lattice of strata contains the empty stratum.

Proof. As observed in [14, Proposition 4.1], a stratum of H s(f) is compact when s ≥ 2 since we can rewrite the s first elementary symmetric polynomials in terms of the s first power sums, and the second power sum defines a sphere.

For the second statement, note that when s = 2, the equations

p 1 ( x ) = c 1  and  p 2 ( x ) = c 2

define a nonempty intersection of a hyperplane, whose normal vector is (1, 1, . . . , 1), and a (d − 1)-sphere. This intersection can either be a (d−2)-sphere in the hyperplane, or it can be one of the two points ±c2d(1,1,,1).

That is, if v = (d) and the stratum Hsv(f) is nonempty, then there can be no other point in Hs(f) since there is no other point in H2(f). So since H s(f) is at least one-dimensional we have H S v ( f ) = , and Hsu(f) contains the empty stratum since v is smaller than any composition u.

For s = 0 and s = 1, the hyperbolic slices will have all the main combinatorial properties we are establishing. But the argument is different from the other cases, so we will restrict s to be at least 2 for now. This allows us to use Lemma 3.2, which will be very helpful. Before we proceed note that we use the convention that the empty set has dimension −1.

Proposition 3.3

If s ≥ 2, then the lattice of strata of H s(f) is graded and the rank of a stratum is one more than its dimension.

Proof. If a stratum Hsu(f) is strictly contained in Hsv(f), then dimHsu(f)<dimHsv(f) by Lemma 3.2 the lattice of strata contains the empty set as its minimal element. Thus any maximal chain in the lattice of strata has length at most dim(Hs(f)) + 1 = ds + 1.

Conversely, suppose that Hsu(f) is strictly contained in Hsv(f), where dimHsu(f)<dimHsv(f)1; we show that there is a stratum Hsw(f) with Hsu(f)Hsw(f)Hsv(f).

If Hsu(f) is empty, Hsv(f) is at least one-dimensional and by Theorem 2.6 its relative interior consists of the polynomials with composition v. But by Lemma 3.2 it is compact and so it must contain a nonempty stratum Hsw(f) in its relative boundary. Then by Lemma 3.2, both Hsv(f) and Hsw(f) contain the empty stratum Hsu(f).

If hHsu(f) and there are no polynomials in Hsu(f) with at least s distinct roots, then Hsu(f)={h} by Theorem 2.7. Since Hsu(f) is zero-dimensional, Hsv(f) is at least two-dimensional, and by Lemma 3.2 and Proposition 2.2 it is compact and contractible. Thus its relative boundary is at least one-dimensional and connected.

If Hsu(f) is not contained in a larger stratum of Hsv(f), it must be an isolated part of the relative boundary of Hsv(f). Since the relative boundary is one-dimensional and connected, Hsu(f) must be the whole boundary. But this is impossible since Hsu(f) is zero-dimensional, thus there is a stratum Hsw(f) which is strictly contained in Hsv(f) and that strictly contains Hsu(f).

Lastly, if there is an hHsu(f) with at least s distinct roots, then we may assume that u = v(h) by Theorem 2.7. Also, we then have that all compositions greater than u occur in H s(f). Since

( u ) s = dim H s u ( f ) < dim H s v ( f ) 1 = ( v ) s 1 ,

we have ℓ(u) < ℓ(v) − 1 and so there is a composition w with u < w < v. By Theorem 2.7, Hsw(f) is of dimension ℓ(w) − s and therefore Hsu(f)Hsw(f)Hsv(f).

Thus any maximal chain will be at least of length ds + 1, and so any maximal chain has length ds + 1. Also, by the above argument any stratum of dimension n ≥ 0 covers a stratum of dimension n − 1, thus its rank must be n + 1. □

Next up is the notion of atomic lattices.

Definition 3.4

In a lattice with smallest element 0, the elements covering 0 are called atoms. The lattice is called atomic if any element can be expressed as the join of atoms.

Lemma 3.5

If s ≥ 2, then for n > 0 any n-dimensional stratum contains at least two distinct (n − 1)-dimensional strata.

Proof. By Proposition 3.3, an n-dimensional stratum Hsu(f) contains an (n − 1)-dimensional stratum Hsv(f). Also, by Lemma 3.2 a nonempty stratum Hsu(f) is compact and so its relative boundary is nonempty but not contractible. But by Proposition 2.2, Hsv(f) is contractible and so it cannot be the whole relative boundary of Hsu(f).

Since Hsv(f) is closed, RelBdHsu(f)\Hsv(f) is relatively open in RelBdHsu(f) with the subspace topology. Thus RelBdHsu(f)\Hsv(f) is (n−1)-dimensional and so there must be another (n−1)-dimensional stratum in Hsu(f).

Proposition 3.6

If s ≥ 2, the lattice of strata of Hs(f) is atomic.

Proof. By convention the empty set is the join of an empty set of atoms, and an atom is naturally the join of itself. Also, by Proposition 3.3, the lattice is graded and a stratum’s rank is its dimension plus one, so the atoms are the zero-dimensional strata.

If Hsu(f) is an n-dimensional stratum and n > 0, then by Lemma 3.5 there are two distinct (n−1)-dimensional strata Hsv(f) and Hsw(f), contained in Hsu(f). Since Hsv(f) and Hsw(f) are distinct, by Proposition 3.3 any stratum that contains both must be at least n-dimensional. Since Hsu(f) is n-dimensional and contains both Hsv(f) and Hsw(f), it must be the join of Hsv(f) and Hsw(f). By induction, both Hsv(f) and Hsw(f) are joins of atoms, and since Hsu(f) is the join of Hsv(f) and Hsw(f) it must also be a join of atoms. □

Lastly, we look at a sort of converse of atomic lattices called coatomic lattices.

Definition 3.7

In a lattice with largest element 1, the elements covered by 1 are called coatoms. The lattice is called coatomic if any element can be expressed as the meet of coatoms.

Lemma 3.8

If s ≥ 2, then for n < ds − 1 any n-dimensional stratum is contained in at least two distinct (n + 1)-dimensional strata.

Proof. Let Hsu(f) be an n-dimensional stratum. If n > 0 then by Theorem 2.7 it is maximal dimensional. Thus there is a polynomial with composition u, and since ℓ(u) − s = n < ds − 1 we have ℓ(u) < d − 1, so u is covered by at least two distinct compositions, say v and w, of length ℓ(u) + 1. So again by Theorem 2.7, Hsv(f) and Hsw(f) must be (n + 1)-dimensional.

If n = 0, then Hsu(f)={h} and since ds > 1, the set Hs(f) is at least two-dimensional. By Proposition 3.3, h is contained in a two-dimensional stratum Hsw(f) and in a one-dimensional stratum Hsv(f)Hsw(f). By Theorem 2.6, Hsv(f) is in the one-dimensional relative boundary of Hsw(f) and h is in the relative boundary of Hsv(f).

According to Corollary 2.8, Hsw(f) is the closure of its relative interior, so there is a connected component C of RelIntHsw(f) such that h is in its closure C. Thus, starting from h, we can traverse its boundary clockwise or counter-clockwise. But since h is one of the relative boundary points of Hsv(f), at most one of the directions consists immediately of polynomials whose composition is v. Thus there must be some other one-dimensional stratum for which h is a boundary point.

Lastly, if n = −1 then Hsu(f) is empty. Since Hs(f) is at least one-dimensional and, by Proposition 3.6, the lattice of strata is atomic, it must contain at least two atoms. Thus the empty stratum is contained in at least two zero-dimensional strata.

Proposition 3.9

If s ≥ 2, the lattice of strata of Hs(f) is coatomic.

Proof. The argument is analogous to the proof of Proposition 3.6, just start the induction from the (ds − 1)-dimensional strata and use Lemma 3.8 instead of Lemma 3.5 for the induction step.

Theorem 3.10

The lattice of strata of H s(f) is graded, atomic and coatomic.

Proof. The only case that remains to check is when s ≤ 1. As we saw in the proof of Theorem 2.7, all compositions occur when s ≤ 1, so the lattice is isomorphic to the lattice of compositions. The lattice of compositions is isomorphic to the face lattice of a (d − 2)-dimensional simplex S. To see this, note that any k + 1 vertices in S determine a k-dimensional face of S.

Similarly there are d − 1 compositions of d of length 2. So if vi1,,vik+1 are k + 1 distinct compositions of length 2, then their first parts are distinct. We may assume that they are ordered such that v1i1<<v1ik+1 and so by the argument in Lemma 1.5 their join is

v i 1 v i k + 1 = v 1 i 1 , v 1 i 2 v 1 i 1 , v 1 i 3 v 1 i 2 , , v 1 i k + 1 v 1 i k , d v 1 i k + 1 ,

which is a composition of length k + 2. Thus any bijection from the set of length 2 compositions to the set of vertices of S induces an isomorphism of lattices. Lastly, by items (i) and (v) of Theorem 2.7 in [19], the face lattice of a simplex is graded, atomic and coatomic.

Similar to the case when s ≤ 1, when s = 2 it can be shown that the lattice of strata is isomorphic to the face lattice of a simplex. Although this does not need to be true in general, as can be seen by intersecting H2(f) from Example 1.4 with the affine hyperplane defined by fixing the third coefficient to be 0. This gives us a lattice which is isomorphic to the face lattice of a quadrilateral.

Thus there are examples for which the lattice of strata is not isomorphic to the face lattice of a simplex; however, it would be interesting to find an answer to the following question for general d and s.

Question 3.11

Is the lattice of strata of a hyperbolic slice polytopal?

We finish this section with an algorithm to compute which compositions occur in Hs(f) based on the compositions of length at most s that occur.

Algorithm 3.12

Suppose that Hs(f) is (ds)-dimensional and that s ≥ 2. Let U denote the set of compositions in Hs(f) of length at most s.

Step 1: Compute the join of every pair of compositions in U:

V : = { u v u , v U a n d u v } .

Step 2: Compute the upward closure of V:

W : = { w v V w i t h v w } .

Then UW is the set of all compositions occurring in H s(f).

Proof. Let wW; then wvu for some v, uU. Thus both v and u occur in Hs(f) and so Hsw(f) contains at least two polynomials. By Theorem 2.7, Hsw(f) is maximal dimensional and therefore there is a polynomial with composition w. Thus all compositions computed in the algorithm occur in Hs(f).

Suppose that a composition u with ℓ(u) > s occurs in Hs(f). Then by Theorem 2.7, Hsu(f) is at least one-dimensional and by Theorem 3.10, Hsu(f) is the join of at least two distinct atoms Hsv(f)={h} and Hsw(f)={g}. We may assume that v and w are the compositions of h and g respectively. Then vwV and vwu, thus uW and it was not left out by the algorithm. □

Remark 3.13

Step 1 in Algorithm 3.12 can be accomplished using the method described in Lemma 1.5. That is, the join of u and v can be computed by first constructing the set

M = u 1 , u 1 + u 2 , , d , v 1 , v 1 + v 2 , , d .

Next, construct the tuple (m1, . . . , mk), where the mi’s are distinct, increasingly ordered and {m1, . . . , mk} = M. Then the join of u and v is the composition

m 1 , m 2 m 1 , m 3 m 2 , , m l m l 1 .

However, compositions and our partial order are both implemented in Sage, see [17], so the algorithm can easily be implemented there.

Remark 3.14

As in [1], [9] and [10] we could have focused on the intersection of Vandermonde varieties and K = {x ∈ ℝd | x1x2 ≤⋅ ⋅⋅ ≤ xd}. That is, we could consider the set

M = x d w 1 x 1 i + w 2 x 2 i + + w d x d i = c i  for all  i [ s ] K

were w1, . . . , wd ∈ ℝ are positive and c1, . . . , cs ∈ ℝ. If xM is of the form x1==xv1<xv1+1== xv1+v2<<xdvl+1==xd we associate to it the composition v(x) = (v1, v2, . . . , vd), and for a composition u we may define a stratum of M as Mu = {yM | v(y) ≤ u}.

When the weights w1, . . . , wd are integers (and by extension rational) then M u is equal to ιuVsu(f)Kl for some monic, univariate hyperbolic polynomial f of degree d (see the proof of Lemma 2.1). However, if the weights are irrational we do not see how to interpret the set M, and so we did not consider such cases. But it can be shown that for any positive real weights, Theorem 2.6 and Theorem 3.10 holds for the poset of strata of M.

As we lack the motivation to consider such cases we do not go into details. But the main difference in the arguments would be to use the map Wl : M → ℝl given by

x j = 1 d w j x j , , j = 1 d w j x j l

instead of using the map PdlΠu in the proof of Theorem 2.6; otherwise the arguments follow through the same way.

4 A short note on concavity

We will discuss some previously established results on the boundary of hyperbolic slices which, if Question 3.11 is answered in the affirmative, would imply that hyperbolic slices always look like partially deflated polytopes. To do this we start by quickly introducing discriminants and subdiscriminants; more information on these objects can be found in [8] and in Chapter 4 of [2].

Let Δd denote the discriminant of a real, monic, univariate polynomial of degree d. If g=td+g1td1++gd ℝ[t] has the roots a1, . . . , ad, the discriminant is defined as

Δ d : = 1 i < j d a i a j 2 .

This is a real polynomial, symmetric in the roots of g, so it may be written in the elementary symmetric polynomials evaluated at (a1, . . . , ad). That is, the discriminant may be written as a polynomial in the coefficients of the univariate polynomial g, and it vanishes whenever the corresponding univariate polynomial has a repeated root.

Let Z(Δd) ⊆ ℝd denote the real algebraic set given by the discriminant. Then the points of Z(Δd) that correspond to polynomials with a repeated real root split the space of real coefficients into ⌊d/2⌋ regions, each of which is characterised by the number of real roots that the polynomials have.

Similarly, we denote by Δd,k the kth subdiscriminant of a real, monic, univariate polynomial of degree d. The subdiscriminant is defined as

Δ d , k : = I [ d ]  and  | I | = d k i , j I  and  i < j a i a j 2

and can also be written as a real polynomial in the coefficients of the univariate polynomial g. We can see that the r first subdiscriminants (including the discriminant) vanish whenever the corresponding univariate polynomial has at most dr distinct roots.

So if h=td+h1td1++hd lies in the real algebraic set Z(Δd,0, . . . , Δd,k) ⊆ ℝd given by the first k subdiscriminants, then h has at most dk distinct roots. If h has exactly dk distinct roots, then it was shown in Proposition 1.3.4 of [12] that a neighbourhood of h in the tangent space of Z(Δd,0, . . . , Δd,k) at h lies in the region which locally about h has the maximal number of real roots. That is, let Th be the tangent space of Z(Δd,0, . . . , Δd,k) at h, where we consider the tangent space as having been translated to h and not centred at the origin. Then there is a neighbourhood N ⊂ ℝd of h such that ThN consists of polynomials that have exactly n real roots and no other polynomial in N has more than n real roots.

Naturally the strata of Hs(f) labelled by compositions of length dk are subsets of Z(Δd,0, . . . , Δd,k). Now suppose that h is hyperbolic and has composition u. Then if the tangent space of Hsu(h) at h is well defined, it must be included in ThAh1Ahs where Ahid is the affine hyperplane defined by fixing the ith coordinate to be equal to hi.

The maximal number of real roots of a polynomial in N is d since h is hyperbolic, thus ThNAh1Ahs Hs(h) and so the tangent space of Hsu(h) at h, intersected with N, must also lie in H s(h). Thus the strata of H s(h) are, in a sense, concave towards Hs(h).

Index of notation

  • [s] = {1, 2, . . . , s}

  • f = t d + f 1 t d 1 + + f d

  • Hs(f) = {h = td + h1td−1 + ⋅⋅ ⋅ + hd | h is hyperbolic and hi = fi for all i ∈ [s]}

  • v(h) denotes the composition of h

  • H s u ( f ) = h H s ( f ) v ( h ) u

  • ℓ(u) denotes the length of the composition u

  • For x ∈ ℝℓ(u) let xu=x1,,x1,,x(u),,x(u), where xi is repeated ui times.

  • ei(x) is the ith elementary symmetric polynomial in d variables.

  • pi(x) is the ith power sum in d variables.

  • V s u ( f ) = x l e 1 x u f 1 = 0 , , ( 1 ) s e s x u f s = 0

  • Kl = {x ∈ ℝl | x1 ≤⋅ ⋅⋅ ≤ xl}

  • For x ∈ ℝl let Πu(x)=tde1xutd1++(1)dedxu

  • Bε(a) is the open ball around a ∈ ℝn of radius ε.

  • uv denotes the join and uv denotes the meet of two elements u and v of a lattice.



Funding statement: This work has been supported by the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie Actions, grant agreement 813211 (POEMA).

Acknowledgements

I am very grateful to Claus Scheiderer for all the helpful discussions, critiques and advice along the way. I am also very grateful to Cordian Riener for bringing to my attention some of the literature on the topic and for greatly simplifying the argument behind Theorem 2.7. Lastly I would also like to thank Tobias Metzlaff whose suggestion led to the pretty picture above.

  1. Communicated by: M. Hering

References

[1] V. I. Arnol’d, Hyperbolic polynomials and Vandermonde mappings. (Russian) Funktsional. Anal. i Prilozhen. 20 (1986), 52–53. English translation in Funct. Anal. Appl. 20 (1986), 125–127. MR847139 Zbl 0608.5800810.1007/BF01077267Suche in Google Scholar

[2] S. Basu, R. Pollack, M.-F. Roy, Algorithms in real algebraic geometry, volume 10 of Algorithms and Computation in Mathematics. Springer 2006. MR2248869 Zbl 1102.1404110.1007/3-540-33099-2Suche in Google Scholar

[3] S. Basu, C. Riener, Vandermonde varieties, mirrored spaces, and the cohomology of symmetric semi-algebraic sets. Found. Comput. Math. 22 (2022), 1395–1462. MR4498438 Zbl 1497.1411710.1007/s10208-021-09519-7Suche in Google Scholar

[4] J. Bochnak, M. Coste, M.-F. Roy, Real algebraic geometry. Springer 1998. MR1659509 Zbl 0912.14023)10.1007/978-3-662-03718-8Suche in Google Scholar

[5] B. Chevallier, Courbes maximales de Harnack et discriminant. In: Séminaire sur la géométrie algébrique réelle, Tome I, II, volume 24 of Publ. Math. Univ. Paris VII, 41–65, Univ. Paris VII, Paris 1986. MR923547 Zbl 0761.14011Suche in Google Scholar

[6] D. A. Cox, J. Little, D. O’Shea, Ideals, varieties, and algorithms. Springer 2015. MR3330490 Zbl 1335.1300110.1007/978-3-319-16721-3Suche in Google Scholar

[7] J.-P. Dedieu, Obreschkoff’s theorem revisited: what convex sets are contained in the set of hyperbolic polynomials? J. Pure Appl. Algebra 81 (1992), 269–278. MR1179101 Zbl 0772.1200210.1016/0022-4049(92)90060-SSuche in Google Scholar

[8] I. M. Gel’fand, M. M. Kapranov, A. V. Zelevinsky, Discriminants, resultants, and multidimensional determinants. Birkhäuser, Boston, MA 1994. MR1264417 Zbl 0827.1403610.1007/978-0-8176-4771-1Suche in Google Scholar

[9] A. B. Givental’, Moments of random variables and the equivariant Morse lemma. (Russian) Uspekhi Mat. Nauk 42 (1987), no. 2(254), 221–222. English translation in Russ. Math. Surv. 42 (1987), No. 2, 275–276. MR898630 Zbl 0647.5702610.1070/RM1987v042n02ABEH001314Suche in Google Scholar

[10] V. P. Kostov, On the geometric properties of Vandermonde’s mapping and on the problem of moments. Proc. Roy. Soc. Edinburgh Sect. A 112 (1989), 203–211. MR1014650 Zbl 0687.5803010.1017/S0308210500018679Suche in Google Scholar

[11] V. P. Kostov, Topics on hyperbolic polynomials in one variable, volume 33 of Panoramas et Synthèses. Société Mathématique de France, Paris 2011. MR2952044 Zbl 1259.12001Suche in Google Scholar

[12] I. Meguerditchian, Géométrie du discriminant réel et des polynômes hyperboliques. PhD thesis, Univ. Rennes 1, 1991.Suche in Google Scholar

[13] I. Meguerditchian, A theorem on the escape from the space of hyperbolic polynomials. Math. Z. 211 (1992), 449–460. MR1190221 Zbl 0763.5800610.1007/BF02571438Suche in Google Scholar

[14] C. Riener, On the degree and half-degree principle for symmetric polynomials. J. Pure Appl. Algebra 216 (2012), 850–856. MR2864859 Zbl 1242.0527210.1016/j.jpaa.2011.08.012Suche in Google Scholar

[15] C. Riener, R. Schabert, Linear slices of hyperbolic polynomials and positivity of symmetric polynomial functions. Preprint 2023, arXiv 2203.08727Suche in Google Scholar

[16] R. P. Stanley, Enumerative combinatorics. Volume 1. Cambridge Univ. Press 2012. MR2868112 Zbl 1247.05003Suche in Google Scholar

[17] The Sage Developers, SageMath, the Sage Mathematics Software System (Version 9.3), 2021. https://www.sagemath.orgSuche in Google Scholar

[18] D. J. Uherka, A. M. Sergott, On the continuous dependence of the roots of a polynomial on its coefficients. Amer. Math. Monthly 84 (1977), 368–370. MR435434 Zbl 0434.3200310.1080/00029890.1977.11994362Suche in Google Scholar

[19] G. M. Ziegler, Lectures on polytopes. Springer 1995. MR1311028 Zbl 0823.5200210.1007/978-1-4613-8431-1Suche in Google Scholar

Received: 2023-09-28
Revised: 2025-02-23
Published Online: 2025-07-19
Published in Print: 2025-07-28

© 2025 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Heruntergeladen am 8.10.2025 von https://www.degruyterbrill.com/document/doi/10.1515/advgeom-2025-0017/html
Button zum nach oben scrollen