1 Introduction

The braid group on n strands can be defined as the fundamental group of the space of monic, complex polynomials in one variable, of degree n and with distinct roots. In this way braids offer a close connection between algebraic geometry and topology. In this paper we illustrate how this connection is beneficial in both directions. We use topological arguments to prove that there exist polynomial maps with singularities and prescribed topological properties and we use insights on polynomial maps to obtain visualizations of certain topological phenomena. Our hope is that we may encourage further research interactions between the two areas.

One well-known intersection of the two areas is the study of links of isolated singularities of polynomial maps. Let \(f:=(f_1,f_2):{\mathbb {R}}^4\rightarrow {\mathbb {R}}^2\) be a real polynomial map in variables \(x_1\), \(x_2\), \(x_3\) and \(x_4\) that satisfies \(f(0,0,0,0)=(0,0)\) and \(\tfrac{\partial f_i}{\partial x_j}(0,0,0,0)=0\) for all \(i=1,2\), \(j=1,2,3,4\). A critical point of f is a point x in \({\mathbb {R}}^4\) where the Jacobian matrix \(\nabla f(x)=\left( \tfrac{\partial f_i}{\partial x_j}(x)\right) _{i,j}\) does not have full rank. We denote the set of critical points of f by \(\Sigma _f\) and the set of zeros of f, i.e., all \(x\in {\mathbb {R}}^4\) with \(f(x)=(0,0)\), by \(V_f\). In particular, we have \((0,0,0,0)\in \Sigma _f\cap V_f\). We say that f has a weakly isolated singularity at the origin if there is a neighborhood U of the origin in \({\mathbb {R}}^4\) such that \(U\cap \Sigma _f\cap V_f=\{(0,0,0,0)\}\). In other words, the rank of \(\nabla f(0,0,0,0)\) should be 0, but for all other \(x\in U\cap V_f\backslash \{(0,0,0,0)\}\) the rank of \(\nabla f(x)\) should be 2.

Let \(S_{\rho }^3\) denote the Euclidean 3-sphere of radius \(\rho \), centered at the origin. If f has a weakly isolated singularity at the origin, then the intersection of \(V_f\) and \(S_{\rho }^3\) results in a closed 1-dimensional submanifold, a link, if \(\rho \) is chosen sufficiently small. Furthermore, the link type of \(L_f:=V_f\cap S_{\rho }^3\), that is, its ambient isotopy class in \(S^3\), is independent of the radius \(\rho \), as long as \(\rho \) is sufficiently small. This link is thus a topological property of the singularity. We call \(L_f\) the link of the singularity.

The name weakly isolated singularity is justified in the sense that it does not impose any restrictions on the link types that arise in this way. Akbulut and King proved that every link is the link of a weakly isolated singularity of some real polynomial map [1]. However, there is also a stronger notion of isolation. We say that the origin is an isolated singularity of f if there is a neighborhood U of the origin in \({\mathbb {R}}^4\) such that \(U\cap \Sigma _f=\{(0,0,0,0)\}\). So the rank of \(\nabla f(0,0,0,0)\) should be 0, but for all other \(x\in U\backslash \{(0,0,0,0)\}\) the rank of \(\nabla f(x)\) should be 2. Naturally, isolated singularities are weakly isolated. A link type is called real algebraic if it arises as the link of an isolated singularity of some real polynomial map.

The definitions above generalize to other dimensions as well as to complex polynomial maps. While links of isolated singularities of complex plane curves \(f:{\mathbb {C}}^2\rightarrow {\mathbb {C}}\) are completely classified, it is not known which links are real algebraic.

A link L in \(S^3\) is called fibered if it is the binding of an open book decomposition of \(S^3\). In other words, there is a fibration map \(\varphi :S^3\backslash L\rightarrow S^1\) with a specified behavior on a tubular neighborhood of L. The term fibration means that \(\varphi \) has no critical points, i.e., no points \(x\in S^3\backslash L\), where the directional derivatives of \(\varphi \) in any three linearly independent directions are all 0. The required behavior of \(\varphi \) on the tubular neighborhood N(L) of L is as follows. Every connected component of N(L) is an open solid torus. Removing L leaves \(S^1\times (D\backslash \{0\})\), where D denotes the open (unit) disk in \({\mathbb {C}}\). On \(S^1\times (D\backslash \{0\})\) we require that \(\varphi (t,z)=\arg (z)\), where \(\arg \) denotes the argument map that sends a nonzero complex number z to \(\tfrac{z}{|z|}\).

The set of fibered links is a promising candidate for the still unknown set of real algebraic links. Milnor proved that all real algebraic links are fibered [28], while Benedetti and Shiota conjecture that the two sets of links are identical [3].

In the last couple of years several constructions of polynomials with isolated singularities have been proposed [7, 10, 13]. These have not resulted in a proof of Benedetti’s and Shiota’s conjecture, but we now have several infinite families of fibered links that are known to be real algebraic. All of these constructions are based on braids and produce so-called semiholomorphic polynomials. This means that when the constructed real polynomial map \(f:{\mathbb {R}}^4\rightarrow {\mathbb {R}}^2\) is written as a mixed polynomial \(f:{\mathbb {C}}^2\rightarrow {\mathbb {C}}\) in complex variables u, v and their complex conjugates \(\bar{u}\) and \(\bar{v}\) (which is clearly possible for any real polynomial map), then f is holomorphic with respect to u, i.e., \(\tfrac{\partial f}{\partial \bar{u}}=0\). From now on we write \(f_u\) for \(\tfrac{\partial f}{\partial u}\).

With a similar construction we obtained the semiholomorphic version of Akbulut’s and King’s result: Every link type arises as the link of a weakly isolated singularity of a semiholomorphic polynomial [12]. One of the main results of this paper is an improvement on this construction. We may perform the construction from [12] to obtain the desired f and obtain any given link \(L_2\) as a sublink of the link of a singularity of \(f_u\), i.e. \(L_2\) is a subset of the connected components of \(L_{f_u}\). We thus have a certain control not only over the topology of \(V_f\) but also over the topology of \(V_{f_u}\) at the same time.

Theorem 1.1

Let \(L_1\) and \(L_2\) be links in \(S^3\). Then there exists a semiholomorphic polynomial \(f:{\mathbb {C}}^2\rightarrow {\mathbb {C}}\) such that both f and \(f_u\) have a weakly isolated singularity at the origin with \(L_1\) as the link of the singularity of f and \(L_2\) a sublink of the link of the singularity of \(f_u\).

A braid on n strands is a set of disjoint parametric curves

$$\begin{aligned} \bigcup _{t\in [0,2\pi ]}\bigcup _{j=1}^n(z_j(t),t)\subset {\mathbb {C}}\times [0,2\pi ], \end{aligned}$$
(1)

where \(z_j(t):[0,2\pi ]\rightarrow {\mathbb {C}}\) are smooth functions with \(z_i(t)\ne z_j(t)\) for all \(t\in [0,2\pi ]\) and all \(i\ne j\). Furthermore, for every index i there should be an index j with \(z_i(0)=z_j(2\pi )\). Since the curves are parametrized by the height coordinate t, no strand can loop back and cross itself. The parametrization induces an orientation on the braid. Writing \(\gamma _j(t)\) for the curve \((z_j(t),t)\), the tangent vector \(\gamma '(t)\) has a positive coefficient of \(\partial _t\), namely \(+1\). We say that the curves are positively transverse to the horizontal planes \({\mathbb {C}}\times \{t\}\) for all \(t\in [0,2\pi ]\). Occasionally, we might encounter a set of parametric curves B in \({\mathbb {C}}\times [0,2\pi ]\) that are transverse to all of these planes, but that are not parametrized by t itself. In this case there is a choice of orientation for B so that is a braid, but it might not match the orientation induced by the parametrization. We could thus consider B as an unoriented braid.

A braid isotopy is an isotopy of a braid in \({\mathbb {C}}\times [0,2\pi ]\) that fixes the start- and endpoints at \(t=0\) and \(t=2\pi \), and maintains the braid property throughout the isotopy. Often a braid isotopy class is also called a braid. When we want to emphasize that we are referring to a representative of an isotopy class instead of the isotopy class itself, we usually call the representative a geometric braid.

Fixing the start- and endpoints \(z_i(0)\) we obtain a group structure on the set of isotopy classes of braids on n strands, where the group operation is given by concatenation and rescaling of the interval. This way we may interpret the braid group on n strands \({\mathbb {B}}_n\) as the fundamental group of the configuration space \(C_n\) of n distinct, unmarked points in the complex plane, where the chosen set of start- and endpoints \(z_i(0)\) corresponds to the basepoint of the loop. \(C_n\) is thus given by the quotient \(({\mathbb {C}}^n\backslash \Delta )/S_n\), where \(\Delta \) is the set of n-tuples where at least two entries are identical and \(S_n\) is the symmetric group on n elements acting on n-tuples by permutation. The space \(({\mathbb {C}}^n\backslash \Delta )\) is an open subset of \({\mathbb {C}}^n\) with the smooth structure of its ambient space \({\mathbb {R}}^{2n}\cong {\mathbb {C}}^n\). By the Quotient Manifold Theorem (e.g., [23]) the quotient space \(C_n\) inherits a unique smooth structure so that the quotient map is a smooth submersion.

By the fundamental theorem of algebra the space \(X_n\) of monic polynomials in one complex variable of degree n and with distinct roots is in bijection with \(C_n\). This correspondence identifies an unordered n-tuple \(\{z_1,z_2,\ldots ,z_n\}\in C_n\) with the polynomial \(g:{\mathbb {C}}\rightarrow {\mathbb {C}}\), \(g(u):=\prod _{j=1}^n(u-z_j)\).

The smooth structure on \(C_n\) thus gives a smooth structure on \(X_n\), so that the bijection established by the fundamental theorem of algebra is a diffeomorphism.

Every geometric braid, parametrized as in Eq. (1), is a loop in \(C_n\) and thus corresponds to a loop in \(X_n\), given by \(g_t:{\mathbb {C}}\rightarrow {\mathbb {C}}\),

$$\begin{aligned} g_t(u)=\prod _{j=1}^n(u-z_j(t)). \end{aligned}$$
(2)

Conversely, the roots of any smooth loop in this space \(X_n\) of polynomials form a geometric braid on n strands.

The natural smooth structure on \(X_n\) explained above matches common intuition as follows. Since the coefficients of a polynomial are smooth functions of its roots (with a smooth local inverse if the roots are distinct), a loop

$$\begin{aligned} g_t(u)=\prod _{j=1}^n (u-z_j(t))=u^n+\sum _{j=0}^{n-1} a_j(t)u^j \end{aligned}$$
(3)

in \(X_n\) is smooth if and only if its coefficients \(a_j(t)\), \(j=1,2,\ldots ,n-1\), are smooth functions of t, which occurs if and only if the strands of the corresponding braid are parametrized by smooth functions \(z_j(t)\), \(j=1,2,\ldots ,n\) [18].

This paper focuses on a subset of this space of polynomials \(X_n\). We define \(\widehat{X}_n\) to be the space of monic polynomials of degree n with distinct roots, distinct critical values and constant term not equal to any of its critical values. If the critical values are distinct, so are the critical points. Since the critical points of a complex polynomial g in one variable u are exactly the roots of \(\tfrac{\partial g}{\partial u}\), which is a polynomial of degree \(n-1\), it follows that we can associate to every loop \(g_t\) in \(\widehat{X}_n\) three braids: one braid on n strands that is formed by the roots of \(g_t\), one braid that is formed by the critical values and one braid on \(n-1\) strands that is formed by the critical points of \(g_t\). The relation between the braid that is formed by the roots of a loop in \(\widehat{X}_n\) with constant term equal to 0, and the braid that is formed by its critical values was studied in detail in [9].

By holomorphicity each critical point of a polynomial p in \(\widehat{X}_n\) must also be a critical point of \(\arg (p)\). Furthermore, it is a saddle point of \(\arg (p)\), that is, it is neither a local maximum nor a local minimum. For this reason we also call the braid that is formed by the critical points of \(g_t\) the saddle point braid of \(g_t\). One major topic of this article is the relation between the braid formed by the roots and the saddle point braid of a given loop of polynomials \(g_t\) in \(\widehat{X}_n\). We prove that any pair of braids can be realized as the roots and saddle point braid of an appropriate loop of polynomials.

Theorem 1.2

Let B and \(B'\) be braids on n and \(n-1\) strands, respectively. Then there is a loop \(g_t:{\mathbb {C}}\rightarrow {\mathbb {C}}\) in \(\widehat{X}_n\) such that

  • the roots \(\{(u,t)\in {\mathbb {C}}\times [0,2\pi ]|g_t(u)=0\}\) form the braid B,

  • the saddle point braid, given by \(\{(u,t)\in {\mathbb {C}}\times [0,2\pi ]|\tfrac{\partial g_t}{\partial u}(u)=0\}\), is \(B'\).

Theorem 1.2 is related to work in [38], where it was shown that not every braid can be realized as the union of roots and saddle point braid of some loop of polynomials, even when certain immediate convexity constraints are taken into account. So by Theorem 1.2 we can control the topology of the braid of roots B and the saddle point braid \(B'\) at the same time, but we do not have control over the braid type of their union, which is a braid on \(2n-1\) strands.

At almost every height \(t\in [0,2\pi ]\) we may order the strands of a braid at that height by the real parts of their complex coordinate. That is, if \(\bigcup _{j=1}^n(z_j(t),t)\), then the first strand at height \(t=t_*\) is the strand with smallest \(\text {Re}(z_j(t_*))\). The second strand has the next smallest value of \(\text {Re}(z_j(t_*))\) and so on. The braid group on n strands is generated by the Artin generators \(\sigma _i\), \(i=1,2,\ldots ,n-1\), which correspond to positive half-twists between the ith strand and the \((i+1)\)st strand.

Definition 1.3

A braid \(B=\prod _{j=1}^{\ell }\sigma _{i_j}^{\varepsilon _j}\) on n strands is called homogeneous if

  1. i)

    for every \(k\in \{1,2,\ldots ,n-1\}\) there is a \(j\in \{1,2,\ldots ,\ell \}\) with \(i_j=k\),

  2. ii)

    for every \(j,j'\in \{1,2,\ldots ,\ell -1\}\), \(i_j=i_{j'}\) implies \(\varepsilon _j=\varepsilon _{j'}\).

In other words, a braid is homogeneous if for all \(i\in \{1,2,\ldots ,n-1\}\) its word contains the generator \(\sigma _i\) if and only if it does not contain its inverse \(\sigma _{i}^{-1}\). In the literature (in particular in [36]) braids are sometimes called homogeneous if they satisfy the second condition above and strictly homogeneous if they satisfy both conditions above.

Since the start- and endpoints of any geometric braid B match, we may identify the \(t=0\)-plane and the \(t=2\pi \)-plane to obtain a link in \({\mathbb {C}}\times S^1\). Embedding this open solid torus as an untwisted neighborhood of a planar circle in \(S^3\) results in a well-defined link in \(S^3\), the closure of the braid B, whose ambient isotopy class in \(S^3\) does not depend on the representative of the braid isotopy class of B.

Stallings showed that closures of homogeneous braids are fibered links [40]. In fact Definition 1.3 has a generalization in the form of T-homogeneous braids, which were introduced and shown to close to fibered links by Rudolph [36]. In Sect. 2 we will review the definition and some properties of T-homogeneous braids.

We say that a geometric braid B parametrized by \(\bigcup _{j=1}^n(z_j(t),t)\) is P-fibered (short for “fibered via polynomials”) if the corresponding loop of polynomials \(g_t(u)=\prod _{j=1}^n(u-z_j(t))\) defines an explicit fibration map via \(\arg (g):({\mathbb {C}}\times S^1)\backslash B\rightarrow S^1\), where \(g:{\mathbb {C}}\times S^1\rightarrow {\mathbb {C}}\), \(g(u,\textrm{e}^{\textrm{i}t})=g_t(u)\). We say that a braid isotopy class is P-fibered if it has a P-fibered representative. Sometimes this condition is weakened to require a P-fibered representative in its conjugacy class in \({\mathbb {B}}_n\). Closures of P-fibered braids are fibered links in \(S^3\) [10], but it is not known if every fibered link is the closure of a P-fibered braid.

T-homogeneous braids are P-fibered [11, 36] and closures of T-homogeneous braids have also recently been proved to be real algebraic [13]. We prove that the loop of polynomials that realizes a T-homogeneous braid as a P-fibered braid can be taken to have a trivial saddle point braid.

Theorem 1.4

Let B be a T-homogeneous braid on n strands. Then there is a loop \(g_t:{\mathbb {C}}\rightarrow {\mathbb {C}}\) in \(\widehat{X}_n\) such that

  • the roots \(\{(u,t)\in {\mathbb {C}}\times [0,2\pi ]|g_t(u)=0\}\) trace out the braid B,

  • \((u,t)\mapsto g_t(u)\mapsto \arg g_t(u)\) is a fibration of \(({\mathbb {C}}\times S^1)\backslash B\) over \(S^1\),

  • the saddle point braid, given by \(\{(u,t)\in {\mathbb {C}}\times [0,2\pi ]|\tfrac{\partial g_t}{\partial u}(u)=0\}\), is the trivial braid on \(n-1\) strands.

The condition that the map \((u,t)\mapsto \arg g_t(u)\) is a fibration means that it does not have any critical points. If the map \((u,t)\mapsto \arg g_t(u)\) for a given loop of polynomials \(g_t\) has a finite number of non-degenerate critical points, i.e., it is a circle-valued Morse map, we call it a pseudo-fibration.

A construction similar to the one employed in [13] results in the following theorem.

Theorem 1.5

Let L be the closure of a T-homogeneous braid B on n strands. Then there is a semiholomorphic polynomial \(f:{\mathbb {C}}^2\rightarrow {\mathbb {C}}\) such that f has an isolated singularity at the origin with link L and \(f_u\) has a weakly isolated singularity at the origin whose link is the unlink with \(n-1\) components.

The structure of the rest of the article is as follows. Section 2 reviews some properties of loops of polynomials in \(\widehat{X}_n\), resulting in a proof of Theorem 1.4 and Theorem 1.2. We also give a new upper bound on the Morse–Novikov number of a link, the minimal number of critical points of a pseudo-fibration. In Sect. 3 we modify the constructions of (weakly) isolated singularities from [12] and [13] to prove Theorem 1.1 and Theorem 1.5. In appendix (Sect. Appendix 1) we then study the argument map \(\arg (g_t)\) of a loop of polynomials \(g_t\) in \(\widehat{X}_n\) and offer different visualizations of the fibration for homogeneous braids as well as of pseudo-fibrations for general braids.

2 Saddle point braids

The space \(\widehat{X}_n\) of monic polynomials of fixed degree n and distinct roots, distinct critical values and constant term different from all critical values is defined in such a way that both the roots and the critical points of a loop in \(\widehat{X}_n\) form a (closed) braid in \({\mathbb {C}}\times S^1\). A parametrization \(z_j(t)\), \(j=1,2,\ldots ,n\), of the roots can be easily obtained by decomposing the polynomials into their irreducible factors, i.e. \(g_t(u)=\prod _{j=1}^n(u-z_j(t))\). Likewise, the critical points of a loop \(g_t\), which form the saddle point braid, are given by \(\bigcup _{j=1}^{n-1}(c_j(t),t)\) where \(\tfrac{\partial g_t}{\partial u}(u)=n\prod _{j=1}^{n-1}(u-c_j(t))\). The relation between these polynomials is then obviously \(g_t(u)=\int _0^u \tfrac{\partial g_t}{\partial u}(w)\textrm{d}w\).

We may write \(v_j(t)=g_t(c_j(t))\), \(j=1,2,\ldots ,n-1\), for the critical values of \(g_t\). The relation between the braid that is parametrized by the \(v_j\) (or more precisely the union of the \(v_j\) and \(\{0\}\times [0,2\pi ]\)) and the braid that is formed by the roots of \(g_t\) was the object of study in [9]. One aspect that continues to be relevant in this paper is that deformations of the braid of critical values lift to deformations in \(\widehat{X}_n\) and thus also to deformations of the braid of roots and the saddle point braid.

More precisely, we may write

$$\begin{aligned} V_n:=\{(v_1,v_2,\ldots ,v_{n-1})\in ({\mathbb {C}}\backslash \{0\})^{n-1}: v_i\ne v_j \text { if }i\ne j\}/S_{n-1}, \end{aligned}$$
(4)

for the space of critical values of polynomials in \(\widehat{X}_n\), where the symmetric group \(S_{n-1}\) acts on \(n-1\)-tuples of non-zero complex numbers by permutation. Note that the critical values are nonzero, since polynomials in \(\widehat{X}_n\) have distinct roots. We then define

$$\begin{aligned} \widehat{V}_n:=\{(v,a_0)\in V_n\times {\mathbb {C}}: a_0\ne v_j \text { for all }j\}. \end{aligned}$$
(5)

Then by Corollary 2.20 in [11] the map \(\theta _n\) that sends a polynomial in \(\widehat{X}_n\) to its set of critical values \(v\in V_n\) and its constant term \(a_0\) is a covering map of degree \(n^{n-1}\).

Suppose now that \(g_t\) is a loop in \(\widehat{X}_n\) and let \(v_t\) denote the loop of the corresponding critical values. That is, if we denote the projection map \(\widehat{V}_n\rightarrow V_n\) by \(\pi \), we have \(v_t=\pi (\theta _n(g_t))\). Any homotopy of \(v_t\) in \(V_n\) is a braid isotopy that therefore extends to an ambient isotopy and gives a homotopy of \(\theta _n(g_t)\) in \(\widehat{V}_n\), which then lifts to a homotopy of \(g_t\) in \(\widehat{X}_n\). By definition of \(\widehat{X}_n\) the braid type of the braid formed by the roots and the braid type of the saddle point braid do not change during this homotopy.

The relation between the braid formed by the roots and the braid formed by the critical values is important in the context of fibrations.

Recall that the property that \(\arg (g)\) is a fibration map means that it does not have any critical points. By [10] \((u_*,t_*)\in ({\mathbb {C}}\times S^1)\backslash B\) is a critical point of \(\arg (g)\) if and only if \(\tfrac{\partial g}{\partial u}(u_*,t_*)=\tfrac{\partial \arg (g)}{\partial t}(u_*,t_*)=0\). This implies that all critical points must lie on the saddle point braid, i.e., there is some \(j\in \{1,2,\ldots ,n-1\}\) such that \(c_j(t_*)=u_*\), where \(c_j(t)\), \(j=1,2,\ldots ,n-1\), parametrize the saddle point braid. The property of being P-fibered is therefore also equivalent to \(\tfrac{\partial \arg (v_j(t))}{\partial t}\ne 0\) for all j and all t, which has the geometric interpretation that each critical value \(v_j(t)\), interpreted as a curve \((v_j(t),t)\) in \({\mathbb {C}}\times [0,2\pi ]\), has a fixed orientation (clockwise or counter-clockwise) with which it twists around \(0\times [0,2\pi ]\subset {\mathbb {C}}\times [0,2\pi ]\).

We now define a particular family of P-fibered braids, the T-homogeneous braids, which feature in Theorems 1.4 and 1.5. They are a generalization of the homogeneous braids defined in Definition 1.3.

Let \(\mathscr {T}\) be an embedded tree graph in the complex plane with n vertices. Furthermore, every edge e of \(\mathscr {T}\) should be associated with a sign \(\varepsilon _e\in \{+,-\}\). See Fig. 1a for an example. After a planar isotopy of \(\mathscr {T}\) we may assume that the vertices of \(\mathscr {T}\) are the n roots of a complex polynomial in \(\widehat{X}_n\). We consider loops in \(\widehat{X}_n\) whose basepoint is given by the n-tuple of the vertices of \(\mathscr {T}\). For every edge e there is a loop \(g_e^{\varepsilon _e}\) in \(\widehat{X}_n\) that exchanges the endpoints of the edge, where the sign of the twist matches \(\varepsilon _e\) as in Fig. 1b. Note that every twist occurs in a neighborhood of the edge. A braid on n strands is called \(\mathscr {T}\)-homogeneous if it has a representative of the form \(g_t=\prod _{j=1}^\ell g_{e_j}^{\varepsilon _{e_j}}\), where for every edge e there is an index \(j\in \{1,2,\ldots ,\ell \}\) with \(e_j=e\). Here \(\prod \) refers to the concatenation of loops, not a product of polynomials. The order in which the twists occur or the number of times each twist occurs is not relevant for this definition. In general we say that a braid B is T-homogeneous if there is some embedded tree \(\mathscr {T}\) with choice of signs \(\varepsilon _e\), such that B is \(\mathscr {T}\)-homogeneous with respect to \(\mathscr {T}\) and the chosen signs. Figure 1c shows an example of a \(\mathscr {T}\)-homogeneous braid for the embedded tree in Fig. 1a. Here and in all figures of braids in this article where a braid is drawn vertically, t is increasing from 0 to \(2\pi \) as we go from the bottom to the top of the picture.

Fig. 1
figure 1

a An embedded tree \(\mathscr {T}\) in the complex plane. The signs \(\varepsilon _e\) are drawn on each edge e. b The loop \(g_e\) exchanges the roots on the endpoints of e as in the upper picture if \(\varepsilon _e=+1\) and as in the lower picture if \(\varepsilon _e=-1\). c A \(\mathscr {T}\)-homogeneous braid with \(\mathscr {T}\) as in Subfigure (a) with t increasing from 0 to \(2\pi \) from the bottom to the top

Every homogeneous braid B is \(\mathscr {T}\)-homogeneous, if \(\mathscr {T}\) is the line graph. Numbering the edges from left to right means that the sign \(\varepsilon _j\) of the edge \(j\in \{1,2,\ldots ,n-1\}\) is the unique sign with which \(\sigma _j\) appears in the homogeneous braid word B.

We now explain how embedded trees arise naturally in the study of complex polynomials. Let \(p:{\mathbb {C}}\rightarrow {\mathbb {C}}\) be a complex polynomial in \(\widehat{X}_n\) with critical values \(v_k\), \(k=1,2,\ldots ,n-1\), such that \(\arg (v_i)\ne \arg (v_j)\) for all \(i\ne j\). Then the argument map \(\arg (p)\) induces a singular foliation on \({\mathbb {C}}\) whose leaves are connected components of level sets of \(\arg (p)\). The singular foliation has two types of singularities; elliptic singularities, which are the roots of p, and hyperbolic singularities, which are the critical points of p. Every singular leaf of this foliation has the shape of a cross, consisting of one line connecting two roots of p and one going to the circle boundary \(\partial D=S^1\) of \({\mathbb {C}}\cong D\). The two lines meet in a critical point of p.

The nth roots of unity divide \(\partial D=S^1\) into n arcs. We denote the arcs by \(A_j\), \(j=1,2,\ldots ,n\), increasing the label as we go around the circle clockwise. The choice of arc that is labeled \(A_1\) is arbitrary, but it should be the same for all polynomials p. Let \(c_k\), \(k=1,2,\ldots ,n-1\) be the critical points of p with \(v_k=p(c_k)\). Since the roots of p are distinct, \(v_k\ne 0\) for all k. We may assume that \(0<\arg (v_1)<\arg (v_2)<\ldots<\arg (v_{n-1})<2\pi \).

We may then associate to each critical point \(c_k\) (or to each critical value \(v_k=p(c_k)\)) a transposition \(\tau _k\in S_n\), where \(S_n\) denotes the permutation group on n elements. As mentioned above, \(c_k\) lies on a unique singular leaf of the singular foliation of D induced by \(\arg (p)\). This singular leaf has two endpoints on \(\partial D=S^1\) that lie on two different arcs \(A_i\) and \(A_j\), \(i\ne j\). Then \(\tau _k=(i\ j)\). It turns out that every such list of ordered transpositions \(\tau _k\) satisfies \(\prod _{k=1}^{n-1}\tau _k=(1\; 2\; \ldots \; n-1 \; n)\). The ordered list \(\{\tau _k\}_{k\in \{1,2,\ldots ,n-1\}}\) is called the cactus of the polynomial p, see [11, 13, 20] for more details. Note that labeling the arcs \(A_j\) clockwise is the convention from [13], which is the opposite of that of [11].

The definition of \(\tau _k\) is illustrated in Fig. 2a. In Fig. 2b we show the singular leaves of the singular foliation induced by \(\arg (p)\), where p is a polynomial of degree 4. The regular leaves can easily be filled in, since they always connect a root to the boundary circle. The cactus of the displayed polynomial is \(\tau _1=(1\ 4)\), \(\tau _2=(2\ 4)\), \(\tau _3=(3\ 4)\).

Fig. 2
figure 2

Defining the cactus of a polynomial p from the singular foliation induced by \(\arg (p)\). a A singular leaf with a critical point \(c_k\) with \(\tau _k=(i\ j)\). b The singular leaves of the singular foliation induced by \(\arg (p)\). c The embedded tree associated with a polynomial p

We may also associate with p in \(\widehat{X}_n\) a planar graph embedded in \({\mathbb {C}}\), whose vertices are the roots of p and whose edges are the parts of the singular leaves that connect two roots, see Fig. 2c. This combinatorial structure is essentially equivalent to the cactus of the polynomial p. The resulting graph is always a tree and up to planar isotopy every embedded tree arises in this way.

Let \(\mathscr {T}\) be the embedded tree associated with a polynomial p in \(\widehat{X}_n\). Suppose that the critical values \(v_j\), \(j=1,2,\ldots ,n-1\), all have distinct arguments. Then the image of \(\mathscr {T}\) under p is a star graph with \(n-1\) edges. Since the vertices of \(\mathscr {T}\) are by definition the roots of p, they all get mapped to the origin in \({\mathbb {C}}\). Every edge of \(\mathscr {T}\) contains a unique critical point \(c_j\) and by definition \(\arg (p)\) is constant along each edge of \(\mathscr {T}\), so that the image of each edge of \(\mathscr {T}\) is a straight line from the origin in \({\mathbb {C}}\) to the corresponding critical value \(v_j=p(c_j)\), see Fig. 3.

Fig. 3
figure 3

Planar tree graph associated with a polynomial p is mapped to a star graph by p

There is a connection between embedded trees and subsets of the BKL-generators \(a_{i,j}\), \(i,j\in \{1,2,\ldots ,n\}\), \(i\ne j\), which were introduced by Birman, Ko and Lee in [5], and which also generate \({\mathbb {B}}_n\). The generator \(a_{i,j}\) represents a positive half-twist between the ith strand and the jth strand such that in the projection the twist lies in front of all strands whose indices are between i and j. See Fig. 4 for an example. In particular, the Artin generator \(\sigma _i\) is given by \(a_{i,i+1}\). BKL-generators (or band generators as they are also called) are often used to describe braided surfaces in \({\mathbb {C}}\times [0,2\pi ]\), where each strand corresponds to a disk and each generator \(a_{i,j}\) to a half-twisted band between the ith and the jth disk. In this way a word in the BKL-generators describes a surface whose boundary is the corresponding braid represented by the same BKL-word.

Fig. 4
figure 4

Band generator \(a_{2,5}\)

For every embedded tree \(\mathscr {T}\) the twists \(g_{e_j}\) around edges \(e_j\), \(j=1,2,\ldots ,n-1\), generate the braid group on n strands, where n is the number of vertices of \(\mathscr {T}\), see [36]. After a planar isotopy of \(\mathscr {T}\) (which does not change the set of corresponding \(\mathscr {T}\)-homogeneous braids up to conjugacy) these generators can be realized as a subset of the band generators or BKL-generators [5, 36]. We call the set \(S_{\mathscr {T}}\) of BKL-generators associated with a given embedded tree \(\mathscr {T}\) the \(\mathscr {T}\)-generators. This means that we can define the set of \(\mathscr {T}\)-homogeneous braids analogously to Definition 1.3. Naturally this definition is equivalent to the one explained above in terms of embedded trees and twists around edges.

Definition 2.1

Let \(S_{\mathscr {T}}=\{s_1,s_2,\ldots ,s_{n-1}\}\) be the set of \(\mathscr {T}\)-generators of \({\mathbb {B}}_n\) for some embedded tree \(\mathscr {T}\) in \({\mathbb {C}}\). Let \(B=\prod _{j=1}^\ell s_{i_j}^{\varepsilon _j}\) be a braid word in the \(\mathscr {T}\)-generators. Then B is a \(\mathscr {T}\)-homogeneous braid word if

  1. i)

    for every \(k\in \{1,2,\ldots ,n-1\}\) there is a \(j\in \{1,2,\ldots ,\ell \}\) with \(i_j=k\),

  2. ii)

    for every \(j,j'\in \{1,2,\ldots ,\ell -1\}\), \(i_j=i_{j'}\) implies \(\varepsilon _j=\varepsilon _{j'}\).

We say that a braid B is T-homogeneous if there exists an embedded tree \(\mathscr {T}\) with \(\mathscr {T}\)-generators \(S_{\mathscr {T}}\) such that B can be represented by a \(\mathscr {T}\)-homogeneous braid word. Thus the set of T-homogeneous braids is the union of all \(\mathscr {T}\)-homogeneous braids, where the union is taken over all embedded trees \(\mathscr {T}\) in \({\mathbb {C}}\).

Clearly, Definition 2.1 reduces to Definition 1.3 if \(\mathscr {T}\) is taken to be the line graph and \(S_{\mathscr {T}}\) is the set of Artin generators.

In [15] we introduced the inhomogeneity \(\beta (B)\) of a braid B, a natural number that measures how far away a given braid word B in Artin generators is from being homogeneous. In particular, \(\beta (B)=0\) if and only if B is a homogeneous braid.

We may now define for every embedded tree \(\mathscr {T}\) the \(\mathscr {T}\)-inhomogeneity \(\beta _{\mathscr {T}}(B)\) of a braid B as follows. Denote the \(\mathscr {T}\)-generators by \(s_1,s_2,\ldots ,s_{n-1}\). Then express B as a word in these generators \(B=\prod _{j=1}^\ell s_{i_j}^{\varepsilon _j}\). Then we count for each \(i\in \{1,2,\ldots ,n-1\}\) the number of sign changes of the generator \(s_i\) as we traverse the braid word cyclically. We add all these numbers and add 2 for every generator that does not appear in the braid word at all, neither with a positive nor with a negative sign. This is expressed as

$$\begin{aligned} \beta _{\mathscr {T}}(B)=&\sum _{i=1}^{n-1}|\{j\in \{1,2,\ldots ,\ell -1\}:\exists k\in \{1,2,\ldots ,\ell -1\} \text { s.t. }i_j=i_{j+k\text { mod }\ell }=i,\nonumber \\&i_{j+m\text { mod }\ell }\ne i \text { for all }m<k\text { and }\varepsilon _j\varepsilon _{j+k\text { mod }\ell }=-1\}|\nonumber \\&+2|\{j\in \{1,2,\ldots ,n-1\}:\text { There is no }k\text { s.t. }i_k=j\}|. \end{aligned}$$
(6)

If \(\mathscr {T}\) is the line graph, the \(\mathscr {T}\)-generators are the usual Artin generators and \(\beta _{\mathscr {T}}(B)=\beta (B)\). Note that by definition \(\beta _{\mathscr {T}}(B)=0\) if and only B is a \(\mathscr {T}\)-homogeneous braid word. We should interpret \(\beta _{\mathscr {T}}(B)\) as a property of a braid word. Of course, we may take any braid, inflate its word artificially by inserting arbitrarily many copies of \(s_js_j^{-1}\) and thereby make \(\beta _{\mathscr {T}}(B)\) arbitrarily large. If we wanted to insist on a topological invariant, we would thus have to take the minimum over all braid words representing the same braid B. For practical purposes, it is of course much simpler to consider \(\beta _{\mathscr {T}}(B)\) as a function from the set of words in \(\mathscr {T}\)-generators (and their inverses) to the natural numbers.

In order to prove Theorem 1.4 we need to find a loop of polynomials \(g_t\) in \(\widehat{X}_n\) whose roots form a given T-homogeneous braid B, realized as a P-fibered geometric braid, and such that its saddle point braid is the trivial braid on \(n-1\) strands. Loops of polynomials that realize B as a P-fibered geometric braid have been constructed in [11, 36] and to some extent (for homogeneous braids) already in [6, 8, 35]. However, these articles do not mention the saddle point braid in this context. We quickly review the main steps in this construction and explain why the corresponding saddle point braid is the trivial braid.

Proposition 2.2

Let \(\mathscr {T}\) be an embedded tree in \({\mathbb {C}}\) with n vertices and let B be a word in the \(\mathscr {T}\)-generators. Then there is a loop \(g_t\) in \(\widehat{X}_n\) such that the roots of \(g_t\) form the braid B, its saddle point braid is the trivial braid on \(n-1\) strands and \(\arg (g):({\mathbb {C}}\times S^1)\backslash B\rightarrow S^1\), \(\arg (g)(u,\textrm{e}^{\textrm{i t}}):=\arg (g_t(u))\), has exactly \(\beta _{\mathscr {T}}(B)\) critical points.

Proof

After a planar isotopy of \(\mathscr {T}\) we may assume that \(\mathscr {T}\) is exactly the embedded tree associated with the polynomial \(p(u):=\prod _{j=1}^n(u-z_j)\) (see, for example, Theorem 5.3 in [11]), where \(z_1,z_2,\ldots ,z_n\) are the vertices of \(\mathscr {T}\). We now study the loop of polynomials in \(\widehat{X}_n\) whose roots form B and that has p as a basepoint.

Each \(\mathscr {T}\)-generator \(s_j\) corresponds to a twist \(g_{e_j}\) along an edge \(e_j\) of \(\mathscr {T}\). Thus B is a concatenation of twists \(g_{e_j}\) along edges \(e_j\) of \(\mathscr {T}\), say \(\prod _{k=1}^\ell g_{e_{j_k}}^{\varepsilon _k}\), where the product refers to concatenation of loops in \(\widehat{X}_n\). Each twist \(g_{e_j}\) corresponds to a very particular motion of the critical values and the constant term in \(\widehat{V}_n\).

Let \(c_1,c_2,\ldots ,c_{n-1}\) be the critical points of p, let \(v_j=p(c_j)\), \(j=1,2,\ldots ,n-1\), be the critical values of p and let \(a_0\) be its constant term. After a small deformation of the embedded graph we may assume that \(\arg (v_i)\ne \arg (v_j)\) if \(i\ne j\). After a translation of \(\mathscr {T}\) in \({\mathbb {C}}\), which does not affect the critical values or the graph structure, we may further assume that one of the \(z_j\) is equal to 0 and so \(a_0=0\ne v_j\) for all j, so in particular, \(p\in \widehat{X}_n\). Since \(\mathscr {T}\) is the embedded tree associated with p, every edge of \(\mathscr {T}\) contains exactly one critical point \(c_j\) of p. We may thus choose the indexing of the edges of \(\mathscr {T}\) such that \(e_j\) contains \(c_j\). There is now a one-to-one correspondence between edges of \(\mathscr {T}\) and critical values of p, where an edge \(e_j\) corresponds to the critical value \(v_j=p(c_j)\), see also Fig. 3. The twist \(g_{e_j}(u)\) can be explicitly realized as a loop in \(\widehat{X}_n\) via \(p(u)-\gamma _j(t)\), where \(\gamma _j(t)\) is a loop in \({\mathbb {C}}\) with basepoint at the origin and the property that it encircles the critical value \(v_j\) counterclockwise in an ellipse that does not contain any other critical values \(v_i\) with \(i\ne j\) as shown in Fig. 5a. We denote the inverse loop of \(\gamma _j(t)\) that encircles \(v_j\) in a clockwise direction by \(\gamma _j(t)^{-1}\). \(\square \)

Fig. 5
figure 5

a Critical values \(v_j\), \(j=1,2,\ldots ,n-1\) (in black), the origin \(0\in {\mathbb {C}}\) (in red) and \(\gamma _j(t)\) (in thick green), both as curves in \({\mathbb {C}}\times S^1\) and as motions in \({\mathbb {C}}\). b Curves \(v_j-\gamma _j(t)\), \(j=1,2,\ldots ,n-1\) (in black), the origin \(0\in {\mathbb {C}}\) (in red) and the constant term \(-\gamma _j(t)\) (in thick blue), both as curves in \({\mathbb {C}}\times S^1\) and as motions in \({\mathbb {C}}\). c A deformation of the curves from Subfigure (b) so that all but one critical value become stationary

Thus we have realized B as the roots of a loop in \(\widehat{X}_n\) that is given by \(\prod _{k=1}^\ell (p-\gamma _{j_k}(t)^{\varepsilon _k})\), where again the product refers to concatenation of loops in \(\widehat{X}_n\). Since the constant term is the only term that depends on t, the saddle point braid of the corresponding loop of polynomials is the trivial braid on \(n-1\) strands.

The motion of the critical values of a loop \(g_{e_j}\) is shown in Fig. 5b. As in Fig. 5c we may deform the loop of critical values and constant term \(-\gamma _j(t)\) such that \(v_i(t)\) does not depend on t if \(i\ne j\) and \(v_j\) encircles \(\{0\}\times [0,2\pi ]\) counterclockwise in an ellipse. Thus the loop of critical values and constant term \(\theta _n(\prod _{k=1}^\ell (p-\gamma _{j_k}(t)^{\varepsilon _k}))\) can be deformed in \(\widehat{V}_n\) so that in each interval \(t\in \left[ \tfrac{2\pi (k-1)}{\ell },\tfrac{2\pi k}{\ell }\right] \) the critical values \(v_i(t)\) with \(i\ne j_k\) do not depend on t, while \(v_{j_k}\) moves on an ellipse around the origin, going counterclockwise if \(\varepsilon _k=1\) and clockwise if \(\varepsilon _k=-1\).

We may now deform the loop of critical values slightly to make \(\partial _t\arg (v_i(t))\) nonzero for all \(t\in \left[ \tfrac{2\pi (k-1)}{\ell },\tfrac{2\pi k}{\ell }\right] \) and all \(i\ne j_k\). All of these deformations are homotopies in \(\widehat{V}_n\) and lift to a homotopy of \(\prod _{k=1}^\ell (p-\gamma _{j_k}(t)^{\varepsilon _k})\) in \(\widehat{X}_n\). The resulting loop \(g_t\) in \(\widehat{X}_n\) therefore still has the same braid of roots and the saddle point braid as \(\prod _{k=1}^\ell (p-\gamma _{j_k}(t)^{\varepsilon _k})\), that is, B and the trivial braid on \(n-1\) strands, respectively.

The corresponding braid of critical values consists of strands that are motions of points on ellipses. We know that critical points of \(\arg (g)\) are exactly points where a critical value \(v_i(t)\) changes its direction with which it moves on its ellipse, clockwise or counterclockwise. Since the direction of motion on the ellipse is given by the signs \(\varepsilon _k\) with \(j_k=i\), this number is exactly

$$\begin{aligned} |\{k\in \{1,2,\ldots ,\ell -1\}:&\exists h\in \{1,2,\ldots ,\ell -1\} \text { s.t. }j_k=j_{k+h\text { mod }\ell }=i,\nonumber \\&j_{k+m\text { mod }\ell }\ne i \text { for all }m<h\text { and }\varepsilon _k\varepsilon _{k+h\text { mod }\ell }=-1\}|. \end{aligned}$$
(7)

If there is no k with \(j_k=i\), i.e., the \(\mathscr {T}\)-generator corresponding to \(g_{e_i}\) does not appear in the braid word with any sign, then \(v_i(t)\) can be taken to be constant. In order to obtain a Morse function, we deform this stationary strand such that \(\arg (v_i(t))\) has exactly two critical points. Thus the number of critical points of \(\arg (g)\) is exactly \(\beta _{\mathscr {T}}(B)\). \(\square \)

Proof of Theorem 1.4

Let B be a \(\mathscr {T}\)-homogeneous braid for some embedded tree \(\mathscr {T}\) with chosen signs. By definition \(\beta _{\mathscr {T}}(B)=0\) and so the theorem follows from Proposition 2.2. \(\square \)

Fibered links in \(S^3\) are exactly the bindings of open book decompositions of \(S^3\). We say that an open book in \(S^3\) is a braided open book if its binding is the closure of a P-fibered braid [11]. The braid axis can then be thought of as a braid axis for the entire open book, not only for the binding, that is, all fiber surfaces (the pages of the open book) are positioned in a very natural way relative to this braid axis. They are all braided surfaces in the sense of Rudolph [36].

An equivalent definition of braided open books in \(S^3\) involves simple branched covers \(S^3\rightarrow S^3\). Montesinos and Morton conjecture that for every fibered link L in \(S^3\) there is a simple branched cover \(\Pi :S^3\rightarrow S^3\) of degree n, branched over a link \(L_{branch}\) such that \(L=\Pi ^{-1}(\alpha )\) is the preimage of some braid axis \(\alpha \) of \(L_{branch}\) and \(L_{branch}\) is the unlink on \(n-1\) components [29].

We showed in [11] that we can construct a simple branched cover \(\Pi :S^3\rightarrow S^3\) from any loop of polynomials \(g_t\), whose roots form a P-fibered geometric braid. The resulting branch link \(L_{branch}\) is exactly the closure of the braid that is formed by the critical values of \(g_t\). The proof of Proposition 2.2 shows that for closures of T-homogeneous braids the conjecture by Montesinos and Morton is true, since the braid of critical values is (exactly like the saddle point braid) the trivial braid on \(n-1\) strands.

In general it is not true that the saddle point braid and the braid of critical values are isotopic, but they must always have the same permutation of strands. So if the saddle point braid is the trivial braid on \(n-1\) strands, then the braid of critical values is a pure braid. That is, every critical value ends at \(t=2\pi \) in the same position where it starts at \(t=0\). Constructing \(\Pi \) from such a loop of polynomials \(g_t\) gives a branch link \(L_{branch}\) with \(n-1\) components, but not necessarily the unlink.

Braided open books have also been studied by Rudolph [36], who calls the saddle point braid without orientation the “derived bibraid”. The name is justified, since the closure of the saddle point braid, as a link in \(S^3\), is transverse to all pages of the open book whose binding is the braid axis of the saddle point braid and also transverse to all pages of the given braided open book. We may therefore choose orientations for the components of the derived bibraid that turn it into a braid (namely, the saddle point braid) relative to its braid axis and another (in general different) choice of orientation turns it into a generalized braid relative to the fibered link L. Singularity theorists might also be interested in Rudolph’s calculation of the Milnor number of a fibered link that is the binding of a braided open book in \(S^3\) in terms of properties of the derived bibraid [36].

The Morse–Novikov number \(\mathscr {M}\hspace{-0.15cm}\mathscr {N}(L)\) of a link L is a natural number that measures how far a given link is from being fibered. In particular, \(\mathscr {M}\hspace{-0.15cm}\mathscr {N}(L)=0\) if and only if L is fibered. In [15] we proved the upper bound \(\mathscr {M}\hspace{-0.15cm}\mathscr {N}(L)\le \beta (B)\) for all links L and all braids B that close to L. We use the discussion above to improve this bound.

The Morse–Novikov number \(\mathscr {M}\hspace{-0.15cm}\mathscr {N}(L)\) is defined to be the minimal number of critical points of any circle-valued Morse map on \(S^3\backslash L\) that displays the usual behavior of an open book in a tubular neighborhood of L, that is, locally it is given by \(\phi :S^1\times (D\backslash \{0\})\rightarrow S^1\), \(\phi (x,z)=\arg (z)\). We will refer to any such map \(\phi \) as a pseudo-fibration. Upper and lower bounds of the Morse–Novikov number in terms of other link invariants have been found in [25, 37] and [33], but we are not aware of any explicit formula or algorithm that computes it.

Corollary 2.3

Let B be a braid on n strands whose closure is the link L. Then \(\mathscr {M}\hspace{-0.15cm}\mathscr {N}(L)\le \min _{\mathscr {T}} \beta _{\mathscr {T}}(B)\), where the minimum is taken over all embedded trees \(\mathscr {T}\) with n vertices.

Proof

From Proposition 2.2 we have for every embedded tree \(\mathscr {T}\) with n vertices a loop \(g_t\) in \(\widehat{X}_n\) whose roots form B and with exactly \(\beta _{\mathscr {T}}(B)\) critical points of \(\arg (g)\). As in [16] we can construct from a Morse function \(\varphi \) on \(S^3\backslash L\), where L is the closure of B with the same number of critical points as follows. By the usual approximation arguments we can assume that the coefficients of \(g_t\) are trigonometric polynomials, i.e., polynomials in \(\textrm{e}^{\textrm{i t}}\) and \(\textrm{e}^{-\textrm{i}t}\). Now take this polynomial expression \(\lambda ^n g_t(\lambda ^{-1} u)\), with \(\lambda \in {\mathbb {R}}\) a parameter and \(n:=\deg _u g_t\), and substitute every \(\textrm{e}^{\textrm{i t}}\) by another complex variable v and every instance of \(\textrm{e}^{-\textrm{i}t}\) by the complex conjugate \(\bar{v}\). We call the resulting semiholomorphic polynomial f. By construction it is \(f|_{v=\textrm{e}^{\textrm{i t}}}=g_t\). It was shown in [16] that for sufficiently small values of \(\lambda \) we have that \(V_f\pitchfork S^3\) is the closure L of B and \(\varphi :=\arg (f)|_{S^3\backslash L}\) has the same number of critical points as g. (The polynomial f does not necessarily have a (weakly) isolated singularity at the origin and the intersection \(S^3_{\rho }\cap V_f\) might be different from L for small radii \(\rho \).) Thus \(\varphi \) satisfies the desired property on a tubular neighborhood of L and we have \(\mathscr {M}\hspace{-0.15cm}\mathscr {N}(L)\le \beta _{\mathscr {T}}(B)\). Since this holds for any embedded tree \(\mathscr {T}\), the result follows. \(\square \)

Note that \(\beta _{\mathscr {T}}(B)\) only depends on the set of \(\mathscr {T}\)-generators, not on the embedded tree \(\mathscr {T}\) per se. Thus the expression \(\min _{\mathscr {T}}\beta _{\mathscr {T}}(B)\) refers to the minimum of a finite set of numbers.

Example 2.4

Consider again the example from Fig. 1c. We already know that it is \(\mathscr {T}_1\)-homogeneous for the embedded tree \(\mathscr {T}_1\) in Fig. 1a. Therefore, \(\beta _{\mathscr {T}_1}(B)=0\) and \(\mathscr {M}\hspace{-0.15cm}\mathscr {N}(L)=0\). However, expressing the same braid in Artin generators gives \(\sigma _2^{-1}\sigma _4\sigma _2^{-1}\sigma _3^{-1}\sigma _1\sigma _2^{-1}\sigma _1^{-1}\sigma _2^{-1}\sigma _4\sigma _3^{-1}\sigma _1\sigma _2^{-1}\sigma _1^{-1}\). We now calculate \(\beta _{\mathscr {T}_2}(B)=\beta (B)\) for the line graph \(\mathscr {T}_2\). Every generator appears with a positive sign or a negative sign, so the last sum in Eq. (6) does not contribute. Furthermore, the generators \(\sigma _2\), \(\sigma _3\) and \(\sigma _4\) all come with a fixed sign. All instances of \(\sigma _2\) and \(\sigma _3\) are negative, while all instances of \(\sigma _4\) are positive. So these strands do not contribute to the count in \(\beta _{\mathscr {T}_2}(B)\). However, the sequence of signs of \(\sigma _1\) as we traverse the braid word reads \(\{+,-,+,-\}\). So there are three sign changes plus one, since the first entry of this list is different from the last one. Thus the bound from [15] would have given \(\mathscr {M}\hspace{-0.15cm}\mathscr {N}(L)=0\le 4=\beta (B)\), while our new improved bound in Corollary 2.3 gives

$$\begin{aligned} 0\le \mathscr {M}\hspace{-0.15cm}\mathscr {N}(L)\le \min _{\mathscr {T}} \beta _{\mathscr {T}}(B)\le \beta _{\mathscr {T}_1}(B)=0 \end{aligned}$$
(8)

and so \(\mathscr {M}\hspace{-0.15cm}\mathscr {N}(L)=\min _{\mathscr {T}} \beta _{\mathscr {T}}(B)=0\).

Proof of Theorem 1.2

Consider a parametrization of the braid \(B'\), say

$$\begin{aligned} \bigcup _{t\in [0,2\pi ]}\bigcup _{j=1}^{n-1}(c_j(t),t)\subset {\mathbb {C}}\times [0,2\pi ], \end{aligned}$$
(9)

with appropriate functions \(c_j:[0,2\pi ]\rightarrow {\mathbb {C}}\). Then \(h_t(u):=n\int _{0}^u\prod _{j=1}^{n-1} (w-c_j(t))\textrm{d}w\) is a loop in the space of monic polynomials of degree n whose critical points form the braid \(B'\) in exactly the given parametrization. After a small deformation of \(h_t\) we may assume that its roots are also distinct. The fact that its critical points form \(B'\) does not change with this small deformation.

Since the roots of \(h_t\) are distinct for all \(t\in [0,2\pi ]\), they form a braid on n strands. However, at this stage we do not know what this braid is. We will call it A. Now apply the construction outlined in the proof to Proposition 2.2 to the braid \(A':=A^{-1}B\) and basepoint \(g_0:=h_0\) to obtain a loop \(g_t\) in \(\widehat{X}_n\) whose roots form the braid \(A'\) and whose saddle point braid is the trivial braid e on \(n-1\) strands. Then the composition of \(h_t\) and \(g_t\) is a loop in \(\widehat{X}_n\) whose roots form the braid \(AA'=B\) and whose critical points form the braid \(B'e=B'\).

3 Singularities of semiholomorphic polynomials

Theorem 1.1 concerns the realization of link types as the link of a weakly isolated singularity of a semiholomorphic polynomial f while at the same time prescribing the link of the singularity of \(f_u\). This type of topological flexibility of the “jet” (in the sense of the literature on the h-principle [21]) of f (while maintaining the link type L) may appear like a typical feature of real algebraic geometry. However, we will see in an example that this flexibility concerning the link of \(f_u\) already appears in the complex setting, which is otherwise known to be very rigid.

First we recall some basic definitions concerning the Newton boundary of a mixed polynomial map \(f:{\mathbb {C}}^2\rightarrow {\mathbb {C}}\) as described in [32]. We may write f as \(f(u,v)=\sum _{i,j,k,\ell \ge 0}c_{i,j,k,\ell }u^i\bar{u}^jv^k\bar{v}^\ell \) with all but finitely many coefficients \(c_{i,j,k,\ell }\) equal to zero. The support of f, denoted by supp(f), consists of all integer lattice points \((\mu ,\nu )\in {\mathbb {Z}}^2\) such that there are non-negative integers \(\mu _1,\mu _2,\nu _1,\nu _2\) with \(\mu =\mu _1+\mu _2\), \(\nu =\nu _1+\nu _2\) and \(c_{\mu _1,\mu _2,\nu _1,\nu _2}\ne 0\).

The Newton polygon of f is the convex hull of \(supp(f)+({\mathbb {R}}^+)^2\). Its boundary consists of a number of vertices and edges. The union of the vertices and compact edges is called the Newton boundary \(\Gamma _f\) of f. There are several properties of \(\Gamma _f\) that can be used to determine if f has a (weakly) isolated singularity, such as convenience and Newton non-degeneracy [32] as well as inner non-degeneracy [2] or partial non-degeneracy [17]. In the case of convenience and Newton non-degeneracy, and inner non-degeneracy it is known that the link of the singularity only depends on the terms of f whose corresponding integer lattice points lie on \(\Gamma _f\). The sum of these terms is called the principal part of f. In other words, adding terms above the Newton boundary does not change the fact that we have a (weakly) isolated singularity and it does not affect the link type. For more details on the Newton boundary of mixed polynomials we point the reader to [2, 17, 32].

Example 3.1

Consider the well-known example of \(f(u,v)=u^p-v^q\) with \(p,q\in {\mathbb {N}}\) with \(p,q>1\), with an isolated singularity at the origin and the (pq)-torus link \(T_{p,q}\) as the link of the singularity. Since f is holomorphic, we may also study \(f_u\) and \(f_v\) with regards to their singularities. However, it is easily seen that neither \(f_u\) nor \(f_v\) has a weakly isolated singularity at the origin.

Instead we may consider \(F(u,v)=f(u,v)-uv^k+vu^{\ell }\) with \(k,\ell \in {\mathbb {N}}\) with \(k\ge q\) and \(\ell \ge p\). First of all, since both added terms lie above the Newton boundary of f and f is convenient and Newton non-degenerate, it follows that F has an isolated singularity at the origin, whose link is again the (pq)-torus link \(T_{p,q}\). But now we have \(F_u(u,v)=pu^{p-1}+v^k\), which has an isolated singularity with link \(T_{(p-1,k)}\), and \(F_v(u,v)=u^{\ell }-qv^{q-1}\), which has an isolated singularity with link \(T_{(\ell ,q-1)}\). We thus have a lot of freedom for the links of \(F_u\) and \(F_v\) without changing the link of F.

This illustrates that the links of singularities of \(F_u\) are not topological invariants of the links of singularities of F. There are different equivalence relations on polynomial map germs with (weakly) isolated singularities related to topological properties (R-equivalence, A-equivalence, V-equivalence, etc.). For example, we might say that two polynomials \(f_1\) and \(f_2\) are A-equivalent if there is a homeomorphism (or diffeomorphism) \(h_1:(B_\varepsilon ^4,0)\rightarrow (B_\varepsilon ^4,0)\) and \(h_2:(B_\varepsilon ^2,0)\rightarrow (B_\varepsilon ^2,0)\) of the 4-ball and 2-ball of radius \(\varepsilon \), respectively, such that on \(B_\varepsilon ^4\) we have \(f_2=h_2\circ f_1\circ h_1\), so that in particular the link of \(f_1\) is ambient isotopic to that of \(f_2\). While the link types of the singularities are topological invariants, that is, they do not depend on the representative of the equivalence class of polynomial maps, we cannot expect the same to be true for the link types of \(f_u\). That is, in general we should expect that the link of \((f_1)_u\) is different from that of \((f_2)_u\) even if \(f_1\) and \(f_2\) are in the same equivalence class. After all, even the property of being semiholomorphic itself depends on the particular variables and therefore changes depending on which representative of the equivalence is under consideration. The links of singularities of first derivatives thus may be used to define much finer equivalence relations that treat the polynomial maps as jets. For this note that there is no particular reason to restrict to semiholomorphic polynomial and derivatives with respect to the complex variable. If \(x_1,x_2,x_3,x_4\) are the real coordinates on \({\mathbb {R}}^4\), we may say that two real polynomial maps \(f_1,f_2:{\mathbb {R}}^4\rightarrow {\mathbb {R}}^2\) are link-equivalent as 1-jets if both have (weakly) isolated singularities with ambient isotopic links and \((f_i)_{x_j}\) has a (weakly) isolated singularity for all \(i=1,2\), \(j=1,2,3,4\), with the link of \((f_1)_{x_j}\) ambient isotopic to the link of \((f_2)_{x_j}\). We might consider such an equivalence relation up to permutation of links of \((f_1)_{x_j}\) so that a simple permutation of the variables does not change the equivalence class. If a polynomial f is semiholomorphic with \(u=x_1+\textrm{i}x_2\), then \(V_{f_u}=V_{f_{x_1}}=V_{f_{x_2}}\), so that the link types of the real derivatives contain the information about the link of the derivative with respect to u.

We now turn our attention to the proofs of Theorem 1.1 and Theorem 1.5. First, we review some techniques to turn loops \(g_t\) of polynomials as in the previous sections into semiholomorphic polynomials with (weakly) isolated singularities.

Since trigonometric polynomials are \(C^1\)-dense in the space of \(2\pi \)-periodic, \(C^1\)-functions, we may approximate any loop of polynomials of fixed degree n by a loop \(g_t\) whose coefficients are polynomials in \(\textrm{e}^{\textrm{i t}}\) and \(\textrm{e}^{-\textrm{i}t}\). Then we may construct a function \(f:{\mathbb {C}}^2\rightarrow {\mathbb {C}}\) with a singularity at the origin via

$$\begin{aligned} f(u,r\textrm{e}^{\textrm{i t}})=r^{kn}g_t\left( \frac{u}{r^k}\right) , \end{aligned}$$
(10)

where k is some sufficiently large natural number. The resulting function f is a holomorphic polynomial with respect to u, but is only a polynomial in the variables \(v=r\textrm{e}^{\textrm{i t}}\) and \(\bar{v}=r\textrm{e}^{-\textrm{i}t}\) if \(g_t\) (and thus its roots as well) satisfy certain symmetry requirements, namely \(g_{t+\pi }=g_t\) or \(g_{t+\pi }(u)=-g_t(-u)\), as shown in [14]. If the roots of \(g_t\) are distinct, then \(\tfrac{\partial g_t}{\partial u}\ne 0\) whenever \(g_t\) is nonzero. Thus \(f_u(u_*,v_*)\ne 0\) , for all \((u_*,v_*)\in V_f\backslash \{0\}\). By the Cauchy–Riemann equations the real Jacobian matrix of f then has full rank at every \((u_*,v_*)\in V_f\backslash \{0\}\). In other words, if the roots of \(g_t\) are distinct, then the singularity at the origin is weakly isolated. With the same techniques as in [7, 15] it can be shown that its link is the closure of the braid formed by the roots of \(g_t\). If the roots of \(g_t\) form a P-fibered geometric braid, then the singularity is isolated [7, 14].

The first author showed in [12] that every link type arises as the link of a weakly isolated singularity of a semiholomorphic polynomial. In [13] it was proved that closures of T-homogeneous braids are real algebraic. Both proofs are constructive and are based on a variation in the idea described above.

In both cases we start with a loop \(g_t\) that satisfies the desired symmetry constraints. However, its roots are not distinct for all \(t\in [0,2\pi ]\), so that they do not form a geometric braid, but a singular braid with intersection points. Defining f from \(g_t\) via Eq. (10) we obtain a semiholomorphic polynomial, whose singularity at the origin is not weakly isolated. By construction f is radially weighted homogeneous, that is, its support in \({\mathbb {Z}}^2\) lies on a straight line of negative slope. In particular, the constructed f is equal to its principal part and is Newton degenerate [17].

We may then find an additional term \(A(v,\bar{v})\) such that \(f+A\) has a (weakly) isolated singularity, whose link is of the desired form, that is, all singular crossings of the singular braids formed by the roots of \(g_t\) are resolved in a very controlled way [12].

The important observation in the context of saddle point braids is that adding \(A(v,\bar{v})\) does not change the derivative with respect to u, i.e., \(f_u=(f+A)_u\).

Proof of Theorem 1.1

The proof is a modification of the construction in [12]. First of all recall that for every link L and every sufficiently large integer n there is a braid on n strands that closes to L. By the same arguments as in [10] we can find for every sufficiently large even integer \(n-1\) a braid B on n strands such that the closure of \(B^2\) contains L as a sublink. Here \(B^2\) denotes the square (or double repeat) of the braid B. It follows that for every pair of links \(L_1\) and \(L_2\) and every sufficiently large even n there exist braids \(B_1\) on n strands and \(B_2\) on \(n-1\) strands, so that \(B_1\) closes to \(L_1\) and the closure of \(B_2^2\) contains \(L_2\) as a sublink.

By Theorem 1.2 there is a loop of polynomials \(h_t\) in \(\widehat{X}_n\) such that the roots of \(h_t\) form the trivial braid and the saddle point braid of \(h_t\) is \(B_2\). Furthermore, after a homotopy of this loop, which does not change the braid types, we may assume that \(h_0\) is a real polynomial, i.e., all of its roots are real numbers. As in [12] we can also construct a loop of polynomials \(g_t\) whose roots form a singular braid \(B_{sing}\) such that for every singular crossing c there is a choice of crossing sign \(\varepsilon _c\in \{\pm 1\}\) that turns \(B_{sing}\) into a classical braid that is isotopic to \(B_1\) if each singular crossing c is replaced by a classical crossing of sign \(\varepsilon _c\). The loop of polynomials \(g_t\) can be taken to consist of real polynomials, so that all of the roots of \(g_t\) are real numbers for all \(t\in [0,2\pi ]\) [12]. We may take the basepoint of \(g_t\) to be \(h_0\).

Consider now the loop that is formed by the composition of \(h_t\) and \(g_t\). We may approximate its coefficients by polynomials in \(\textrm{e}^{\textrm{i t}}\) and \(\textrm{e}^{-\textrm{i}t}\). Call the resulting loop of polynomials \(G_t\). The approximation can be chosen arbitrarily \(C^1\)-close and we can interpolate the original coefficient functions, so that the roots of \(G_t\) form the singular braid \(B_{sing}\) and the corresponding saddle point braid is \(B_2\). Note that the saddle point braid of \(g_t\) was the trivial braid, since it is a real polynomial whose n roots have at most multiplicity 2.

Therefore, the roots of the loop \(G_{2t}\) form the singular braid \(B_{sing}^2\) and its saddle point braid is \(B_2^2\). Furthermore, the loop obviously satisfies the symmetry constraint \(G_{2(t+\pi )}=G_{2t}\), so that f defined from \(G_{2t}\) as in Eq. (10) is a radially weighted homogeneous semiholomorphic polynomial.

As in [12] we may now find an extra additive term \(A(v,\bar{v})\), which gives \(f+A\) a weakly isolated singularity at the origin and resolves all singular crossings in a way that guarantees that the resulting link is the closure of \(B_1\), that is, \(L_1\). The only difference to the proof in [12] is that \(G_{2t}\) is not a loop of real polynomials. However, recall that all singular crossings of the roots of \(G_t\) occur in intervals where \(G_{2t}\) is an arbitrarily close approximation of \(g_t\), which is a loop of real polynomials. This is sufficient for the arguments from [12], so that \(f+A\) has a weakly isolated singularity whose link is \(L_1\). Furthermore, as observed above, we have \((f+A)_u=f_u\), so that the zeros of \((f+A)_u\) are exactly \(\bigcup _{j=1}^n(r^kc_j(t),r \textrm{e}^{\textrm{i t}})\subset {\mathbb {C}}^{2}\) with \(r\ge 0\), \(t\in [0,2\pi ]\) and \(\bigcup _{j=1}^n(c_j(t),t)\) a parametrization of the saddle point braid of \(G_{2t}\). Since the critical points of \(G_{2t}\) are all distinct (they form a braid), this implies that all roots of \((f+A)_u\) except the origin are regular points of \((f+A)_u\). This means that \((f+A)_u\) has a weakly isolated singularity at the origin and its link is the closure of \(B_2^2\), which by construction contains \(L_2\) as a sublink. \(\square \)

We see from the proof that in general the link of \((f+A)_u\) has extra components besides \(L_2\). This is because by construction, the saddle point braid of \(G_{2t}\) is a 2-periodic braid \(B_2^2\). If on the on the other hand \(L_2\) is the closure of a 2-periodic braid \(B_2^2\) on \(n-1\) strands, such that n is at least the braid index of \(L_1\), then the link of the singularity of \((f+A)_u\) is exactly \(L_2\) without any extra components. Take, for example, the figure-eight knot, which is the closure of the 2-periodic braid \((\sigma _1\sigma _2^{-1})^{2}\) on three strands. So we may take \(B_2=\sigma _1\sigma _2^{-1}\) and \(B_1\) any braid on four strands. Then the construction above gives a semiholomorphic polynomial map \(f:{\mathbb {C}}^2\rightarrow {\mathbb {C}}\) with a weakly isolated singularity at the origin and the closure of \(B_1\) as the link of the singularity. Furthermore, the derivative \(f_u\) also has a weakly isolated singularity at the origin and its link is the figure-eight knot. So in this case the given knot (\(4_1\)) is not only a sublink of the link of the singularity but equal to the link of the singularity. No additional components are required.

Proof of Theorem 1.5

The theorem follows almost immediately from the construction in [13], where we start with a loop \(g_t\), whose roots form a P-fibered geometric braid whose closure is the unknot and that satisfies \(g_{t+\pi }(u)=-g_t(-u)\). We deform the critical values of \(g_t\), so that there are some values of t for which there is a critical value equal to 0, which means that at that value of t the roots of the deformed \(g_t\) are not disjoint and thus form a singular braid as t varies from 0 to \(2\pi \). It is proved in [13] that this deformation can be done in such a way that the resulting critical values \(v_j(t)\) still satisfy the condition \(\tfrac{\partial \arg (v_j(t))}{\partial t}\ne 0\) for all \(j=1,2,\ldots ,n-1\) and all values of t with \(v_j(t)\ne 0\). Furthermore, we may assume that the coefficients of the loop of polynomials \(\widehat{g}_t\), which is defined to be the endpoint of the lift of the deformation of the loop of critical values, are polynomials in \(\textrm{e}^{\textrm{i t}}\) and \(\textrm{e}^{-\textrm{i}t}\).

By Theorem 1.4 we can do this procedure, starting with a loop \(g_t\) whose saddle point braid is the trivial braid on \(n-1\) strands. Since the critical values are distinct throughout the deformation, the same is true for the critical points. It follows that the saddle point braid of \(\widehat{g}_t\) is also the trivial braid on \(n-1\) strands.

We may then construct f from \(\widehat{g}\) as in Eq. (10). By [13] there is a polynomial \(A(v,\bar{v})\) such that \(f+A\) has an isolated singularity at the origin and its link is the closure of the given T-homogeneous braid. Again, adding A does not affect \(f_u\), so that the roots of \((f+A)_u\) are precisely \((r^kc_j(t),r\textrm{e}^{\textrm{i}t})\subset {\mathbb {C}}^2\), where \(r\ge 0\), \(t\in [0,2\pi ]\) and \(\bigcup _{j=1}^{n-1}(c_j(t),t)\) is a parametrization of the saddle point braid of \(\widehat{g}_t\), which is the trivial braid on \(n-1\) strands. Thus \((f+A)_u\) has a weakly isolated singularity at the origin, whose link is the unlink on \(n-1\) components.