October 02, 2025
We study continuous approximate solutions to polynomial equations over the ring \(C(X)\) of continuous complex-valued functions over a compact Hausdorff space \(X\). We show that when \(X\) is one-dimensional, the existence of such approximate solutions is governed by the behaviour of maps from the fundamental pro-group of \(X\) into braid groups.
Let \(X\) be a compact Hausdorff space, and denote by \(C(X)\) the C*-algebra of continuous complex-valued functions on \(X\). For each positive integer \(n\) and each \(n\)-tuple of continuous functions \(a_0,a_1, \ldots, a_{n-1} \in C(X)\), we obtain a monic polynomial over the ring \(C(X)\) of degree \(n\) given by \[\begin{align} P(x, z) = z^n + a_{n-1}(x)z^{n-1} + \cdots + a_1(x)z + a_0(x) \end{align}\] One obvious question one can ask is whether \(P(x,z)\) has any roots in the ring \(C(X)\), that is, if there exist functions \(f \in C(X)\) such that \(P(x,f(x)) = 0\) for all \(x \in X\).
The problem of characterizing those topological spaces \(X\) for which every monic polynomial \(P(x,z)\) over \(C(X)\) has a global continuous solution has been extensively studied [1]–[4]. Work of Gorin and Lin [5] has also established necessary and sufficient conditions for \(P(x,z)\) to split into linear factors over \(C(X)\) in the case that \(P(x_0,z)\) has no repeated roots for each \(x_0 \in X\).
While there has been considerable effort in understanding polynomials over \(C(X)\), most of the results in the literature have been algebraic in nature, concerning only properties \(C(X)\) has as a ring. However, in [6], Miura takes an analytic approach, making use of the sup-norm on \(C(X)\). Together with Kawamura in [2], this analytic approach is further investigated and for continua of dimension at most one, they obtain a topological characterization for when \(C(X)\) possesses approximate \(n^{\text{th}}\) roots.
Recently, these analytic results have found a fruitful application in the continuous model theory of Banach algebras (following [7]). Indeed, in [8] it was shown that the approximate algebraic closure of \(C(X)\) is an \(\forall\exists\)-axiomatizable property in the language of unital C*-algebras and an explicit axiomatization is given. On the other hand, it is also shown that the property of (exact) algebraic closure of \(C(X)\) is not axiomatizable in this language.
In this paper, we provide topological characterizations for when \(C(X)\) is approximately algebraically closed, in the case that \(X\) has covering dimension one. The most general result concerns a shape-theoretic invariant, the fundamental pro-group \(\underline{\pi}_1(X)\) of a space \(X\) (see Section 4 for a review of shape theory), and maps from \(\underline{\pi}_1(X)\) into the braid group on \(n\) strands \(\mathcal{B}_n\).
Theorem 1 (Theorem 33). Suppose that \(X\) is a continuum with covering dimension at most one. Then \(C(X)\) is approximately algebraically closed if and only if for every \(n \geq 1\), the image of any morphism \(\underline{\pi}_1(X) \to \mathcal{B}_n\) lies in the subgroup of pure braids, i.e. in the kernel of the canonical map \(\mathcal{B}_n \to S_n\).
Using this criterion, we are able to provide examples of continua that are not (exactly) algebraically closed, but are approximately algebraically closed (Examples 38). An important application of our results is to the study of co-existentially closed continua. This class of continua was first introduced by Bankston in [9], and are known to be one-dimensional [10], so our results apply (Example 1). We believe that this is an important first step towards a classification of co-existentially closed continua. From the perspective of continuous model theory, the next step towards this goal would include studying more general classes of equations over one-dimensional spaces, namely, non-monic equations, multivariable equations, and *-polynomials (i.e. polynomials in \(z\) and \(\overline{z}\)).
The remainder of this paper is organized as follows. In Section 2, we set up and discuss some elementary facts concerning polynomials over \(C(X)\). We also review some relevant facts about braid groups. Section 3 is analytic in nature, as we relate approximate roots of an arbitrary polynomial to exact roots of better-behaved polynomials (i.e. those with non-vanishing discriminant). In Section 4, we review some material on shape theory we will need, especially results about the fundamental pro-group, which is the primary topological invariant we use in this paper. In Section 5, we formulate how exact roots of well-behaved polynomials are related to morphisms from the fundamental pro-group into braid groups. In Section 6, we combine the results in the previous sections to obtain Theorem 1 above, and provide examples of how to use it. Finally, in Section 7 we explain some simpler characterizations for obtaining approximate roots of low-degree polynomials, and give some examples to show that these simpler invariants cannot be used for higher degree polynomials.
We are grateful to Christopher Eagle and George Elliott for helpful discussions and comments.
This paper is concerned with monic polynomials over the ring of continuous functions on a compact Hausdorff space (i.e. a compactum). Given a compactum \(X\), we write a monic polynomial \(P \in C(X)[z]\) of degree \(n \geq 1\) as \[\begin{align} P(x,z) = z^{n} + a_{n-1}(x)z^{n-1} + \cdots + a_1(x)z + a_0(x) \end{align}\] where \(a_0, a_1, \ldots, a_{n-1} \in C(X)\). We typically denote polynomials over \(C(X)\) using capital letters such as \(P\) and \(Q\), whereas we typically denote polynomials over \(\mathbb{C}\) with lowercase letters such as \(p\) and \(q\). Observe that if \(C(X,\mathbb{C}^n)\) denotes the C*-algebra of continuous functions from \(X\) to \(\mathbb{C}^n\), then we have a natural identification of vector spaces \[\begin{align} \{ P \in C(X)[z] : P \text{ is monic with degree n}\} &\longleftrightarrow C(X,\mathbb{C}^n) \\ z^{n} + a_{n-1}(x)z^{n-1} + \cdots + a_1(x)z + a_0(x) &\longleftrightarrow (a_0, a_1, \ldots, a_{n-1}) \end{align}\] This identification, endows the space of monic polynomials of degree \(n\) over \(C(X)\) with the structure of a Banach space, and in particular this space is topologised with the norm \[\begin{align} \left\Vert P \right\Vert = \max_{0 \leq i < n} \left\Vert a_i \right\Vert_{\infty} \end{align}\] for \(P(x,z) = z^{n} + a_{n-1}(x)z^{n-1} + \cdots + a_1(x)z + a_0(x)\). Due to this identification, we make no distinction between a monic polynomial of degree \(n\) in \(C(X)[z]\) and a continuous map \(X \to \mathbb{C}^n\).
Observe that for each \(x_0 \in X\) we have a natural evaluation map \(\operatorname{ev}_{x_0} : C(X,\mathbb{C}^n) \to \mathbb{C}[z]\) given by \[\begin{align} \operatorname{ev}_{x_0}(a_0, a_1, \ldots, a_{n-1}) = z^n +a_{n-1}(x_0)z^{n-1} + \cdots + a_{1}(x_0)z + a_0(x_0). \end{align}\] For a monic polynomial \(P(x,z) \in C(X)[z]\), we denote the monic polynomial \(\operatorname{ev}_{x_0}(P) \in \mathbb{C}[z]\) by \(P(x_0,z)\). Applying the usual evaluation map \(\operatorname{ev}_{z_0} : \mathbb{C}[z] \to \mathbb{C}\) at a point \(z_0 \in \mathbb{C}\), we obtain, for each \(x_0 \in X\) and each \(z_0 \in \mathbb{C}\), a complex number \(P(x_0,z_0)\) given by \(P(x_0,z_0) = \operatorname{ev}_{z_0} \operatorname{ev}_{x_0}(P)\).
Above, the discussion focuses solely on a monic polynomial in \(C(X)[z]\) in terms of its coefficients, but there is an alternative and useful perspective of thinking in terms of the roots. This is done as follows: To each monic polynomial \(P \in C(X)[z]\) of degree \(n\), we can associate a map \(X \to \mathbb{C}^n/S_n\) defined by associating to each \(x_0 \in X\), the multiset of roots of of the complex polynomial \(P(x_0,z) \in \mathbb{C}[z]\) (this is well-defined because the polynomial is monic and hence does not drop degree as \(x_0\) varies). By a consequence of Rouché’s Theorem (see Lemma 7), this map is continuous.
An especially important class of polynomials are those that never have repeated roots, in which case this map descends to a continuous map \(X \to \operatorname{UConf}_n(\mathbb{C})\), the unordered configuration space on \(n\) points in \(\mathbb{C}\). This space of polynomials with no repeated roots also has a description in terms of the coefficients as follows. Let \(\Delta : \mathbb{C}^n \to \mathbb{C}\) denote the monic discriminant function, i.e. \(\Delta(a_0, a_1, \ldots, a_{n-1})\) is the discriminant of the degree \(n\) monic polynomial \(z^n + a_{n-1}z^{n-1} + \cdots + a_1 z + a_0\). Then \(\Delta\) is some polynomial in \(a_0, a_1, \ldots, a_{n-1}\), and its zero-set defines the (monic) discriminant variety \[V(\Delta) = \{ (a_0, a_1, \ldots, a_{n-1}) \in \mathbb{C}^n : \Delta(a_0, a_1, \ldots, a_{n-1}) = 0 \}\] which is a closed subset of \(\mathbb{C}^n\). Then the open subset \(B_n := \mathbb{C}^n \setminus V(\Delta)\) of \(\mathbb{C}^n\) is the space of coefficients for monic polynomials of degree \(n\) having no repeated roots, so the space of degree \(n\) monic polynomials over \(C(X)\) having no repeated roots is identified with the subset \(C(X,B_n)\) in \(C(X, \mathbb{C}^n)\).
The space of monic polynomials over \(\mathbb{C}\) exhibits a natural \(\mathbb{C}^{\times}\)-action given by scaling the roots. Precisely, if we denote this action by \(\star\), then a number \(\mu \in \mathbb{C}^\times\) acts on a polynomial \(\prod_{1\leq i \leq n}(z - \lambda_j) \in \mathbb{C}[z]\) by \[\begin{align} \mu \star \prod_{1\leq i \leq n}(z - \lambda_j) = \prod_{1 \leq i \leq n}(z-\mu \lambda_j) \end{align}\] In terms of the coefficients, this action is given by \[\begin{align} \mu \star (z^n + a_{n-1}z^{n-1} + \cdots + a_1 z + a_0) = z^n + \mu a_{n-1} z^{n-1} + \cdots + \mu^{n-1}a_1 z + \mu^n a_0 \end{align}\] The perspective of this action in terms of the roots shows that the action clearly restricts to the space \(B_n\). An important observation is that with respect to this action, the discriminant function \(\Delta : B_n \to \mathbb{C}^\times\) is homogeneous of degree \(n(n-1)\): \[\Delta(\mu \star p) = \prod_{1 \leq i < j \leq n} \mu^2(\lambda_i - \lambda_j)^2 = \mu^{n(n-1)} \Delta(p)\] This implies that \(\Delta : B_n \to \mathbb{C}^\times\) is a fibre bundle for \(n \geq 2\). Indeed, if \(n\geq 2\), then we can find neighbourhoods \(U\) and \(V\) of \(1 \in \mathbb{C}^\times\) such that the map \(\mu \mapsto \mu^{n(n - 1)}\) defines a homeomorphism from \(U\) to \(V\). Then given a polynomial \(p_0 \in B_n\) and its discriminant \(\delta_0 = \Delta(p_0)\) we have the trivialization map \[\begin{align} U \times \Delta^{-1}(\delta_0) & \longrightarrow \Delta^{-1}(V\delta_0) \\ (\mu, p) & \longmapsto \mu \star p \end{align}\] which is a homoemorphism as the inverse is given by taking some \(q \in \Delta^{-1}(V\delta_0)\) to \((\mu, \mu^{-1} \star p)\) for \(\mu = (\Delta(q)/\delta_0)^{1/n(n-1)}\), where the \(n(n - 1)\)th root is taken using the inverse \(V \to U\) we specified above. This shows that \(\Delta: B_n \to \mathbb{C}^\times\) is a fibre bundle. We define the subset \(B_n' = \Delta^{-1}(1) \subseteq B_n\) to be the fibre over \(1 \in \mathbb{C}^\times\), i.e. the space of monic, degree \(n\) polynomials with no repeated roots having discriminant equal to \(1\). This subset \(B_n'\) is path-connected ([5]).
Let \(X\) be a compactum. As mentioned in the introduction, the question of when a monic polynomial \(P \in C(X)[z]\) has a root in the ring \(C(X)\) (i.e. when there is a continuous function \(f \in C(X)\) such that \(P(x,f(x)) = 0\) for all \(x\in X\)) has been extensively studied. If \(f \in C(X)\) is a root of \(P\), we sometimes say \(f\) is an exact root to distinguish it from the approximate roots that are treated below. If \(P\) factors completely over \(C(X)\), i.e. when there exist \(f_1, \ldots, f_{\deg P} \in C(X)\) such that \(P(x,z) = \prod_{k=1}^{\deg P} (z-f_k(x))\), we say that \(P\) is completely solvable. Complete solvability of polynomials over \(C(X)\) was studied by Gorin and Lin in [5], and we adapt many of their methods here.
Definition 2. For \(X\) a compactum, we say that the ring \(C(X)\) is algebraically closed (resp. completely solvable) if every non-constant monic polynomial over \(C(X)\) has an exact root (resp. is completely solvable).
In this paper, we are also interested in approximate roots of monic polynomials over \(C(X)\). Let us make this notion precise.
Definition 3. Suppose that \(P(x,z)\) is a monic polynomial over \(C(X)\), and let \(\varepsilon> 0\) be given. We say that a function \(f \in C(X)\) is an \(\varepsilon\)-approximate root of \(P\) provided that \(|P(x,f(x))| < \varepsilon\) for all \(x \in X\). We say that \(P\) has approximate roots if \(P\) has an \(\varepsilon\)-approximate root for all \(\varepsilon> 0\).
Definition 4. Let \(X\) be a compactum. We say that \(C(X)\) is approximately algebraically closed if every non-constant monic polynomial over \(C(X)\) has approximate roots.
In the case when a polynomial \(P\) of degree \(n \geq 1\) has no repeated roots, there is a topological description for when \(P\) admits a continuous root that lends itself well to homotopical methods. Following [5], we define the solution space for a monic polynomial of degree \(n\) with no repeated roots to be the space \[E_n = \{((a_0, a_1, \ldots, a_{n-1}), z_0) \in B_n \times \mathbb{C}: \operatorname{ev}_{z_0}(a_0, a_1, \ldots, a_{n-1}) = 0\} \subset \mathbb{C}^n \times \mathbb{C}\] which consists of elements in \(B_n\) along with a choice of root (such a choice always exists by the Fundamental Theorem of Algebra). Then we have the natural projection map \(E_n \to B_n\) that forgets the root, and we see that given a polynomial \(P: X \to B_n\) a solution of \(P\) is nothing but a continuous lift \(\lambda: X \to E_n\) making the diagram \[\begin{tikzcd} & {E_n} \\ X & {B_n} \arrow[from=1-2, to=2-2] \arrow["\lambda", dashed, from=2-1, to=1-2] \arrow["P"', from=2-1, to=2-2] \end{tikzcd}\] commute. The projection map \(E_n \to B_n\) is an \(n\)-sheeted covering map, and so understanding the fundamental group of these spaces will be crucial for determining when we can find lifts (and hence roots of polynomials). As explained in [5], the fundamental group of \(B_n\) is the Artin braid group on \(n\)-strands, which we denote by \(\mathcal{B}_n\). Moreover, \(B_n\) is an Eilenberg-MacLane space \(K(\mathcal{B}_n, 1)\) [11]
The fundamental group \(M_n\) of \(E_n\) is then an index \(n\) subgroup of \(\pi_1(B_n) \cong \mathcal{B}_n\) and can be identified with the subgroup of braids that fix the first strand. Note that the way this group \(M_n\) sits inside \(E_n\) depends on a choice of basepoint. Choosing a basepoint \(b_0 \in B_n\), we have \(n\) choices for a basepoint of \(E_n\) that maps to \(b_0\), corresponding to the \(n\) distinct choices of roots for the monic polynomial \(b_0\); let’s denote them \(z_1, \dots, z_n\). Then the start and end points of the strands for a braid in \(\pi_1(B_n, b_0) \cong \mathcal{B}_n\) are labelled by \(z_1, \dots, z_n\) and the subgroup \(M_{n, i} = \pi_1(E_n, z_i)\) of \(\mathcal{B}_n\) consits of the braids whose strand starting at \(z_i\) also ends at \(z_i\).
Another important subgroup of \(\mathcal{B}_n\) is \(N_n := \bigcap_i M_{n,i}\), which corresponds to the subgroup consisting of braids with each of its \(n\) strands starting and ending at the same point (this is the kernel of the natural homomorphism \(\tau: \mathcal{B}_n \to S_n\) into the symmetric group). This subgroup corresponds to the \(n!\)-sheeted covering of \(B_n\) consisting of polynomials together with an ordering of their respective \(n\) roots.
A final important subgroup of \(\mathcal{B}_n\) is the commutator subgroup \(\mathcal{B}_n'\). It turns out that \(\mathcal{B}_n/\mathcal{B}_n' \cong \mathbb{Z}\) and that the discriminant function \(\Delta: B_n \to \mathbb{C}^\times\) induces the abelianization map \(\pi_1(\Delta): \mathcal{B}_n \to \mathbb{Z}\). Then using the fibre sequence \[B_n' \longrightarrow B_n \longrightarrow \mathbb{C}^\times\] that we found in Section 2.2, we get the exact sequence \[\begin{tikzcd} {0 \cong \pi_2(\mathbb{C}^\times)} & {\pi_1(B_n')} & {\pi_1(B_n)} & {\pi_1(\mathbb{C}^\times) \cong \mathbb{Z}} & {\pi_0(B_n') \cong 0} \arrow[from=1-1, to=1-2] \arrow[from=1-2, to=1-3] \arrow["{\pi_1(\Delta)}", from=1-3, to=1-4] \arrow[from=1-4, to=1-5] \end{tikzcd}\] which gives us that \(\pi_1(B_n') \cong \mathcal{B}_n'\) (justifying the notation) and that the inclusion map on spaces \(B_n' \to B_n\) induces the inclusion map on subgroups \(\mathcal{B}_n' \cong \pi_1(B_n') \to \pi_1(B_n) \cong \mathcal{B}_n\).
The goal of this section is to reduce the study of approximate roots for all polynomials over \(C(X)\) to the study of exact roots of polynomials over \(C(X)\) that have no repeated roots. This will allow us to later use the methods in the paper [5], which treats the no repeated roots case. Thus, the main theorem of this section is the following:
Theorem 5. Let \(X\) be a compactum with \(\dim X \leq 1\) and let \(n\) be a positive integer. The following are equivalent:
Every monic polynomial of degree \(n\) with coefficients in \(C(X)\) has approximate roots;
Every monic polynomial of degree \(n\) with coefficients in \(C(X)\) with no repeated roots has approximate roots;
Every monic polynomial of degree \(n\) with coefficients in \(C(X)\) with no repeated roots has an exact root.
The basic strategy is as follows. First, to prove that (2) implies (3) in the above, we need to understand under what circumstances having roots is a stable property, i.e. when a given polynomial over \(C(X)\) having approximate roots actually implies that it has an exact root. Then, to show that (3) implies (1), we are lead to investigate when we can perturb the coefficients of a polynomial so that it has no repeated roots, and this is where the restriction on the dimension of \(X\) appears. The key ingredient is the simple observation that the roots of a polynomial are bounded by the coefficients. In our case, this means the following.
Lemma 6. Let \(X\) be a compactum, let \(P(x,z) = z^n + a_{n-1}(x)z^{n-1} + \cdots + a_1(x)z + a_0(x)\) be a polynomial with coefficients in \(C(X)\), and let \(M := \max\{ \left\Vert a_0 \right\Vert, \left\Vert a_1 \right\Vert, \ldots, \left\Vert a_{n-1} \right\Vert \}\). If \(f \in C(X)\) is an \(\varepsilon\)-approximate solution of \(P\), then \(\left\Vert f \right\Vert_{\infty} \leq 1 + \varepsilon+ M\).
Proof. Let \(f \in C(X)\) be an \(\varepsilon\)-approximate solution of \(P\). Take any \(x_0 \in X\), and let \(w_0 := P(x_0,f(x_0))\). Since \(f\) is an \(\varepsilon\)-approximate solution of \(P\), we know that \(|w_0| < \varepsilon\). Observe that \[\begin{align} z^n + a_{n-1}(x_0)z^{n-1} + \cdots + a_1(x_0)z + (a_0(x_0) - w_0) \end{align}\] is a polynomial over \(\mathbb{C}\) possessing \(f(x_0)\) as a root. Hence by Cauchy’s bound for the root of a polynomial, we have that \[\begin{align} |f(x_0)| &\leq 1 + \max\{ |a_0(x_0) - w_0|, |a_1(x_0)| , \ldots, |a_{n-1}(x_0)| \} \\ &\leq 1 + \max\{ |a_0(x_0)| + |w_0|, |a_1(x_0)| , \ldots, |a_{n-1}(x_0)| \} \\ &\leq 1 + |w_0| + \max\{ |a_0(x_0)|, |a_1(x_0)| , \ldots, |a_{n-1}(x_0)| \} \\ &\leq 1 + \varepsilon+ \max\{ |a_0(x_0)|, |a_1(x_0)| , \ldots, |a_{n-1}(x_0)| \} \\ &\leq 1 + \varepsilon+ M \end{align}\] As the above bound holds for any \(x_0 \in X\), we obtain \(\left\Vert f \right\Vert_{\infty} \leq 1 + \varepsilon+ M\). ◻
As mentioned in Section 2.1, we can think of any monic polynomial \(P(x,z)\) over \(C(X)\) in terms of its roots, by considering the map \(X \to \mathbb{C}^n/S_n\) sending some \(x_0 \in X\) to the multiset of roots of \(P(x_0,z) \in \mathbb{C}[z]\). We take some time to formulate what this means precisely, and show that this map is continuous.
Lemma 7. The roots of a polynomial depend continuously on its coefficients. Precisely, given any monic polynomial \(p(z) = z^n + a_{n - 1}z^{n - 1} + \cdots + a_1z + a_0\) that factors as \[p(z) = \prod_{i = 1}^r (z - \lambda_i)^{m_i}\] for some distinct roots \(\lambda_i \in \mathbb{C}\) with respective multiplicities \(m_i\), and any \(\delta \in (0, \frac{1}{2}\min_{i \neq j} |\lambda_i - \lambda_j|)\), there exists an \(\varepsilon> 0\) so that any monic polynomial \(q(z) = z^n + b_{n - 1}z^{n - 1} + \cdots + b_1z + b_0\) with coefficients satisfying \(\max_{0\leq k <n} |a_k - b_k| < \varepsilon\) has exactly \(m_i\) roots counted with multiplicity in the disk \(B_{\delta}(\lambda_i) = \{z \in \mathbb{C}: |z - \lambda_i| < \delta\}\).
Proof. For each \(0 \leq i \leq r\) consider the circle \(C_i = \{z \in \mathbb{C}: |z - \lambda_i| = \delta\}\), which by construction contains no root of \(p(z)\). Let \[\mu = \min_{0 \leq i \leq r} \min_{z \in C_i} |p(z)|\] where the minima are achieved and positive by compactness of the circles and continuity of \(p\). Then since the values of a polynomial over \(\mathbb{C}\) depend continuously on its coefficients, we can find an \(\varepsilon> 0\) such that if \(q(z) = z^n + b_{n - 1}z^{n - 1} + \cdots + b_1z + b_0\) is a polynomial with \(|a_k - b_k| < \varepsilon\) then \[\max_{0 \leq i \leq r} \max_{z \in C_i} |q(z) - p(z)| < \mu\] But then for any such \(q(z)\), we have \(|q(z) - p(z)| < |p(z)|\) for all \(z \in C_i\), which by Rouché’s Theorem implies that \(p(z)\) and \(q(z)\) have the same number of zeroes in the disk \(\{z \in \mathbb{C} : |z - \lambda_i| < \delta\}\) as required. ◻
We can reformulate the above lemma as follows. Given a monic polynomial \(p \in \mathbb{C}[z]\), let us denote by \(\sigma_p\) the multiset of roots of \(p\). Here, we view \(\sigma_p\) as an element of the quotient space \(\mathbb{C}^n/S_n\), which we endow with the quotient topology. In these terms, the above lemma becomes the following.
Corollary 8. The map \[\begin{align} \sigma: \mathbb{C}^n & \longrightarrow \mathbb{C}^n/S_n \\ (a_0, \dots, a_{n - 1}) & \longmapsto \sigma_{z^n + a_{n - 1}z^{n - 1} + \cdots + a_1z + a_0} \end{align}\] taking the coefficients of a polynomial in \(\mathbb{C}[z]\) to its multiset of roots is continuous. In particular, given a monic polynomial \(P(x,z)\) with coefficients in \(C(X)\), the map \(X \to \mathbb{C}^n/S_n , x_0 \mapsto \sigma_{P(x_0,z)}\) is continuous.
For each complex number \(z \in \mathbb{C}\) and each multiset \(A \in \mathbb{C}^n/S_n\), we define the distance from \(z\) to \(A\) to be the number \(d(z,A) := \min_{w \in A} |z-w|\). This assignment \[\begin{align} \mathbb{C}\times (\mathbb{C}^n/S_n) & \longrightarrow [0, \infty) \\ (z, A) & \longmapsto d(z, A) \end{align}\] is a continuous function. Using these facts, we now show that approximate roots of \(P(x,z)\) are uniformly close to the multiset of roots of \(P(x,z)\). This is reminiscent of the notion of a weakly stable predicate in the sense of [12].
Lemma 9. Let \(X\) be a compactum, and let \(P(x,z) = z^n + a_{n-1}(x)z^{n-1} + \cdots + a_1(x)z + a_0(x)\) be a monic polynomial with coefficients in \(C(X)\). Then given any \(\delta > 0\) there exists an \(\varepsilon> 0\) such that all \(\varepsilon\)-approximate solutions \(f \in C(X)\) of \(P\) are within \(\delta\) of one of the roots of \(P(x, z)\), i.e. \(d(f(x), \sigma_{P(x,z)}) < \delta\) for all \(x \in X\).
Proof. Let \(R := 2 + \max_{0 \leq i < n} \left\Vert a_i \right\Vert\) and consider the set \[Y := \{(x, w) \in X \times \mathbb{C}\;: \;|w| \leq R, \quad d(w, \sigma_{P(x,z)}) \geq \delta\}\] Using Corollary 8 we see that the function \((x, w) \mapsto d(w, \sigma_{P(x,z)})\) is continuous, and hence \(Y\) is a closed set. Being a subset of the compact \(X \times \{w \in \mathbb{C} : |w| \leq R\}\), we see that \(Y\) is also compact. Let \[\varepsilon:= \min (\{ |P(x,w)| : (x,w) \in Y \} \cup \{1\})\] which is positive as \(Y\) does not contain a pair \((x, w)\) such that \(P(x,w) = 0\).
Now assume that \(f \in C(X)\) is an \(\varepsilon\)-approximate solution of \(P\). Then by Lemma 6 we get that \(|f(x)| \leq 1 + \varepsilon+ \max_{0 \leq i < n} \left\Vert a_i \right\Vert \leq R\) for all \(x \in X\). However for each \(x_0 \in X\), we have that \(|P(x_0,f(x_0))| < \varepsilon\), which by our choice of \(\varepsilon\), implies that \(|P(x_0,f(x_0))| < |P(x,w)|\) for all \((x,w) \in Y\). But this means that for each \(x_0 \in X\), the pair \((x_0, f(x_0))\) cannot be in \(Y\), and since \(|f(x_0)| \leq R\) this is only possible when \(d(f(x_0), \sigma_{P(x_0,z)}) < \delta\). ◻
This allows us to prove our first stability result: for a polynomial with no repeated roots, having approximate roots is equivalent to having an exact root.
Theorem 10. Let \(X\) be a compactum. If \(P(x,z)\) is a monic polynomial with coefficients in \(C(X)\) which has no repeated roots, then \(P\) has approximate roots if and only if it has an exact root.
Proof. Clearly if \(P\) has an exact root then it has approximate roots. Conversely assume that \(P\) has approximate roots. Let \(\delta : X \to [0,\infty)\) be the function which assigns to each \(x \in X\), the minimum distance between any two roots of of the complex polynomial \(P(x,z)\), i.e. \[\begin{align} \delta(x) = \min\{ |\lambda - \mu| : \lambda, \mu \in \sigma_{P(x,z)}, \lambda \neq \mu \} \end{align}\] Observe that \(\delta\) is non-vanishing since \(P\) has no repeated roots, and it is a continuous function by Lemma 7. In particular, it achieves a minimum since \(X\) is compact; let \(\alpha := \frac{1}{2}\min_{x \in X} \delta(x)\). By Lemma 9, we can find an \(\varepsilon> 0\) such that for all \(\varepsilon\)-approximate solutions \(f\) of \(P\), we have that \(d(f(x),\sigma_{P(x,z)}) < \frac{\alpha}{3}\) for all \(x \in X\).
Now since \(P\) has approximate roots, we can find an \(\varepsilon\)-approximate solution \(f \in C(X)\) of \(P\); by above we have that \(d(f(x), \sigma_{P(x,z)}) < \frac{\alpha}{3}\) for all \(x \in X\). This means that for each \(x \in X\), there exists some \(\lambda_x \in \sigma_{P(x,z)}\) such that \(|f(x) - \lambda_x| < \frac{\alpha}{3}\); in fact, this \(\lambda_x\) is unique, since if \(\mu_x \in \sigma_{P(x,z)}\) is any other root of \(P(x,z)\), then \[\begin{align} |f(x) - \mu_x| \geq |\lambda_x - \mu_x| - |f(x) - \lambda_x| \geq \delta(x) - \frac{\alpha}{3} \geq 2\alpha - \frac{\alpha}{3} \geq \frac{5}{3}\alpha \end{align}\] Thus we obtain a well-defined function \(g : X \to \mathbb{C}\) given by \(g(x) = \lambda_x\). By definition we have that \(P(x,g(x)) = 0\) for all \(x \in X\), so to show that \(P\) has an exact root, it suffices to show that \(g\) is continuous. To this end, take any \(x_0 \in X\) and assume that \(\varepsilon' > 0\) is given. Since \(f\) is continuous, and by Corollary 8, the function \(X \to \mathbb{C}^n/S_n , x \mapsto \sigma_{P(x,z)}\) is continuous and \(g(x_0) \in \sigma_{P(x_0,z)}\), we can find an open subset \(U\) of \(x_0\) such that whenever \(x \in U\), we have that \(d(g(x_0), \sigma_{P(x,z)}) < \min(\frac{\alpha}{3}, \varepsilon')\) and \(|f(x) - f(x_0)| < \frac{\alpha}{3}\). But this implies that for all \(x \in U\), there is some \(\lambda \in \sigma_{P(x,z)}\) such that \(|\lambda - g(x_0)| < \min(\frac{\alpha}{3}, \varepsilon')\), which implies that \[\begin{align} |\lambda - f(x)| &\leq |\lambda - g(x_0)| + |g(x_0) - f(x_0)| + |f(x_0) - f(x)| \\ &\leq \frac{\alpha}{3} + \frac{\alpha}{3} + \frac{\alpha}{3} = \alpha \end{align}\] The calculation \((\clubsuit)\) above shows that if \(\mu\) is any root of \(P(x,z)\) which is not \(g(x)\), then it must be that \(|f(x) - \mu| \geq \frac{5}{3}\alpha\). But since \(|\lambda - f(x)| \leq \alpha\) and \(\lambda \in \sigma_{P(x,z)}\), this means that \(\lambda = g(x)\). Thus \(|g(x) - g(x_0)| = |\lambda - g(x_0)| < \varepsilon'\). As this holds for all \(x \in U\), this proves that \(g\) is continuous, as desired. ◻
In particular, the notion of having approximate roots is interesting only for polynomials which can have repeated roots. Instead of placing a condition on the kind of polynomials we consider, we can instead ask the question of what kind of spaces \(X\) have the property that every monic polynomial over \(C(X)\) having approximate roots actually has exact roots. Our second stability result is the following theorem, which shows that the main topological obstruction to passing from approximate solutions to exact solutions is local-connectedness. This is a generalization of [6], where it is shown that if \(X\) is locally connected, then \(C(X)\) is square-root closed if and only if it has approximate square roots.
Theorem 11. Let \(X\) be a locally connected compactum. If \(P(x,z)\) is a monic polynomial with coefficients in \(C(X)\), then \(P\) has approximate roots if and only if it has an exact root.
Proof. Obviously if \(P(x,z)\) has an exact root, then it has approximate roots. Conversely assume that \(P(x,z)\) has approximate roots. Then for each integer \(k \geq 1\), we can find a \(\frac{1}{k}\)-approximate root \(f_k \in C(X)\) of \(P(x,z)\). If we can show that \(\{f_k\}_{k = 1}^\infty\) has a convergent subsequence \(\{f_{k_{l}}\}_{l=1}^{\infty}\) converging to some \(f_\infty \in C(X)\), then passing to this subsequence we get \[|P(x, f_\infty(x))| = \lim_{l \to \infty} |P(x, f_{k_l}(x))| \leq \lim_{k \to \infty} \frac{1}{k} = 0\] for each \(x \in X\), showing that \(f_\infty\) is an exact solution. To this end, we will show that \(\{f_k\}_{k = 1}^\infty\) has a convergent subsequence by checking that this sequence is uniformly bounded and equicontinuous, and then applying the Arzelà–Ascoli Theorem.
By Lemma 6, we have that \(\left\Vert f_k \right\Vert_\infty \leq 2 + M\) for all \(k\), showing the sequence is uniformly bounded. For equicontinuity, pick a point \(x_0 \in X\) and \(\varepsilon> 0\). Then we may factor our polynomial at this point \[P(x_0, z) = \prod_{i = 1}^r (z - \lambda_i)^{m_i}\] for some distinct roots \(\lambda_i \in \mathbb{C}\) with respective multiplicities \(m_i\). Let \(\varepsilon_0 := \frac{1}{2}\min(\varepsilon, \min_{i \neq j} |\lambda_i - \lambda_j|)\). By Lemma 7 we can find a neighbourhood \(U_1 \subset X\) of \(x_0\) such that \[\begin{align} \bigcup_{x \in U_1} \{ w \in \mathbb{C} : P(x,w) = 0 \} \subset \bigcup_{i=1}^{r} B_{\frac{\varepsilon_0}{4}}(\lambda_i) \end{align}\] and since \(X\) is locally connected, we can find a connected open neighbourhood \(U\) of \(x_0\) with \(U \subset U_1\). Now by Lemma 9, we can find a positive integer \(k_0\) such that \(d(f_k(x), \sigma_{P(x,z)}) < \frac{\varepsilon_0}{4}\) for all \(k \geq k_0\) and all \(x \in X\). Hence if \(x \in U\) and \(k \geq k_0\), then we can find a root \(w_{k,x}\) of \(P(x,z)\) such that \(|f_k(x) - w_{k,x}| < \frac{\varepsilon_0}{4}\), and an \(i_{k,x} \in \{1, \ldots, r\}\) such that \(w_{k,x} \in B_{\frac{\varepsilon_0}{4}}(\lambda_{i_{k,x}})\). The triangle inequality immediately implies \(|f_k(x) - \lambda_{i_{k,x}}| < \frac{\varepsilon_0}{4} + \frac{\varepsilon_0}{4} = \frac{\varepsilon_0}{2}\) for all \(k\geq k_0\) and all \(x \in U\). It follows that \(f_k(x) \in \bigcup_{i = 1}^r B_{\frac{\varepsilon_0}{2}}(\lambda_i)\) for all \(x \in U\) and \(k \geq k_0\). However, since \(\varepsilon_0 < \min_{i \neq j} |\lambda_i - \lambda_j|\), it follows that the family of disks \(\{B_{\frac{\varepsilon_0}{2}}(\lambda_i)\}_{i=1}^{r}\) are disjoint, so by connectedness of \(U\) it follows that for each \(k \geq k_0\), \(f_k(U)\) must lie in a single such disk; that is, for each \(k \geq k_0\) we can find some \(i_k\) such that \(f_k(U) \subset B_{\frac{\varepsilon_0}{2}}(\lambda_{i_k})\). Therefore for all \(x \in U\) and all \(k \geq k_0\) we have that: \[\begin{align} |f_k(x) - f_k(x_0)| \leq |f_k(x) - \lambda_{i_k}| + |\lambda_{i_k} - f_k(x_0)| < \frac{\varepsilon_0}{2} + \frac{\varepsilon_0}{2} < \varepsilon \end{align}\] Therefore the sequence \(\{f_k\}_{k=k_0}^{\infty}\) is equicontinuous, and since \(f_1, \ldots, f_{k_0}\) are continuous functions, it follows that the sequence \(\{f_k\}_{k=1}^{\infty}\) is equicontinuous as desired. ◻
Corollary 12. If \(X\) is a locally connected compactum, then \(C(X)\) is approximately algebraically closed if and only if it is algebraically closed.
We now turn our attention to perturbation, i.e. describing when we can perturb a polynomial so that it has no repeated roots. As noted in Section 2.1, the space of polynomials with no repeated roots is identified with the subset \(C(X, B_n)\) of the C*-algebra \(C(X,\mathbb{C}^n)\), so precisely this means describing when \(C(X,B_n)\) is dense in \(C(X,\mathbb{C}^n)\).
In order to do this, it will be helpful to introduce the following notation: For each subset \(M\) of \(\mathbb{C}^n\), we let \(\mathcal{A}(X,M)\) denote the set of continuous functions \(f : X \to \mathbb{C}^n\) whose image avoids \(M\), i.e. \[\mathcal{A}(X,M) := \{ f : X \to \mathbb{C}^n : \operatorname{im}(f) \cap M = \emptyset \}\] Observe that by definition, \(C(X,B_n) = \mathcal{A}(X,V(\Delta))\) where \(V(\Delta)\) is the discriminant variety. The following lemma describes some elementary properties of these sets.
Lemma 13. Let \(X\) be a compactum and \(n\) a positive integer.
(i) If \(\{ M_{\alpha} \}_{\alpha \in I}\) is a family of subsets of \(\mathbb{C}^n\), then \(\mathcal{A}\left(X, \bigcup_{\alpha \in I} M_{\alpha}\right) = \bigcap_{\alpha \in I} \mathcal{A}(X,M_{\alpha})\).
(ii) If \(M\) is a closed subset of \(\mathbb{C}^n\), then \(\mathcal{A}(X,M)\) is an open subset of \(C(X,\mathbb{C}^n)\).
Proof. Part (i) is clear. For part (ii), let \(M\) be a closed subset of \(\mathbb{C}^n\) and assume that \(f \in \mathcal{A}(X,M)\). Since \(X\) is compact, \(f(X)\) is a compact subset of \(\mathbb{C}^n\) not intersecting \(M\); hence \(\delta := \operatorname{dist}(f(X), M) > 0\). Then if \(\left\Vert g - f \right\Vert < \frac{\delta}{2}\), it follows that for all \(x \in X\) and all \(m \in M\), we have that \[\begin{align} |g(x) - m| \geq |f(x) - m| - |g(x) - f(x)| \geq \delta - \frac{\delta}{2} = \frac{\delta}{2} \end{align}\] Taking the infimum over all \(x \in X\) and all \(m \in M\) we obtain that \(\operatorname{dist}(g(X), M) \geq \frac{\delta}{2} > 0\); therefore \(g(X) \cap M = \emptyset\), so \(g \in \mathcal{A}(X,M)\). Thus \(B_{\frac{\delta}{2}}(f) \subset \mathcal{A}(X,M)\). This proves that \(\mathcal{A}(X,M)\) is open. ◻
The basic idea that we will show is that if \(\dim X \leq 1\) and if \(M \subset \mathbb{C}^n\) has real codimension at least \(2\), then the avoiding set \(\mathcal{A}(X,M)\) is dense in \(C(X,\mathbb{C}^n)\). The reason for the restriction on the dimension is that we suspect that our proof below generalizes to the statement that if \(M \subset \mathbb{C}^n\) has real codimension at least \(1 + \dim X\), then \(\mathcal{A}(X,M)\) is dense in \(C(X,\mathbb{C}^n)\). However, since we are mostly concerned with the complex hypersurface \(V(\Delta)\), we restrict our attention to the case when \(M\) has real codimension two, and thus when \(\dim X \leq 1\). As a starting point, we prove this for the complex hypersurface \(M\) cut out by the equation \(z_1z_2 \cdots z_n = 0\).
Lemma 14. Let \(X\) be a compactum with \(\dim X \leq 1\), and let \(n\) be a positive integer. We denote by \(M \subset \mathbb{C}^n\) the complex hypersurface defined by \(z_1z_2\cdots z_n = 0\).
(i) \(\mathcal{A}(X,M)\) is open and dense in \(C(X,\mathbb{C}^n)\).
(ii) If \(h : \mathbb{C}^n \to \mathbb{C}^n\) is a homeomorphism, then \(\mathcal{A}(X,h(M))\) is open and dense in \(C(X,\mathbb{C}^n)\).
Proof.
(i) Take any \(f \in C(X,\mathbb{C}^n)\), and let \(\varepsilon> 0\). Denote by \(f_1,\ldots, f_n\) the components of \(f\). Then \(f_j \in C(X)\) for all \(1 \leq j \leq n\). Since \(\dim X \leq 1\), the invertible elements form a dense subset of \(C(X)\) [13], so we can find non-vanishing functions \(g_1, \ldots, g_n\) in \(C(X)\) such that \(\left\Vert f_j - g_j \right\Vert_{\infty} < \varepsilon\) for each \(1 \leq j \leq n\). Let \(g = (g_1,\ldots, g_n)\). Then \(g \in C(X,\mathbb{C}^n)\), and for each \(x \in X\) we have that \(g_1(x)g_2(x) \cdots g_n(x) \neq 0\), so \(g(x) \notin M\). Therefore \(g \in \mathcal{A}(X,M)\). Moreover, \[\begin{align} \left\Vert f-g \right\Vert = \max_{1 \leq j \leq n} \left\Vert f_j - g_j \right\Vert_{\infty} < \varepsilon \end{align}\] Therefore \(\mathcal{A}(X,M)\) is dense in \(C(X,\mathbb{C}^n)\). Finally, to see that \(\mathcal{A}(X,M)\) is open, just observe that \(M\) is a closed subset of \(\mathbb{C}^n\) and apply Lemma 13.
(ii) Since \(X\) is compact, \(C(X,\mathbb{C}^n)\) has the compact-open topology, so it follows that the the pushforward \(h_* : C(X,\mathbb{C}^n) \to C(X,\mathbb{C}^n)\) is a homeomorphism. By part (i), \(\mathcal{A}(X,M)\) is open and dense in \(C(X,\mathbb{C}^n)\), so it follows that \(h_*(\mathcal{A}(X,M))\) is open and dense in \(C(X,\mathbb{C}^n)\). Then using that \(h\) is bijective, it is easy to see that \(h_*(\mathcal{A}(X,M)) = \mathcal{A}(X,h(M))\), proving (ii).
◻
We now use the above Lemma to show \(\mathcal{A}(X,M)\) is dense in \(C(X,\mathbb{C}^n)\) whenever \(M\) is a smooth submanifold of \(\mathbb{C}^n\) with real codimension at least two. For that, we will need to following way of loosely “covering" such a submanifold with complex hyperplanes.
Lemma 15. Let \(N \subset \mathbb{C}^n\) denote a smooth submanifold of \(\mathbb{C}^n\) with real codimension at least \(2\). Then there exists a countable sequence \((f_k)_{k = 1}^{\infty}\) of autodiffeomorphisms of \(\mathbb{C}^n\) such that \[N \subset \bigcup_{k = 1}^{\infty} f_k(\{z \in \mathbb{C}^n : z_n = 0\}).\]
Proof. Here we identify \(\mathbb{C}^n\) with \(\mathbb{R}^{2n}\), and we let \(j\) denote the real dimension of \(N\), which we know is at most \(2n-2\).
Assume first that \(N\) is contained in the graph of a smooth function \(\varphi : \mathbb{R}^j \to \mathbb{R}^{2n - j}\), meaning that \(N \subseteq\{(x, y) \in \mathbb{R}^j \times \mathbb{R}^{2n - j} : y = \varphi(x)\}\) (where by permuting the coordinates we can assume the first \(j\) coordinates are the independent ones). Then letting \(f: \mathbb{R}^{2n} \to \mathbb{R}^{2n}\) be defined by \(f(x, y) = (x, y + \varphi(x))\), which is a diffeomorphism with inverse is given by \((x, y) \mapsto (x, y - \varphi(x))\), we have \[\begin{align} N \subset f(\{(x, y) \in \mathbb{R}^j \times \mathbb{R}^{2n - j} : y = 0\}). \end{align}\] Since the zero set of these last \(2n - j\) coordinates is contained in the zero set of the last two coordinates (which is \(\{z \in \mathbb{C}^n : z_n = 0\}\)), we have completed the proof in this case.
For a general \(N \subseteq\mathbb{R}^{2n}\) we just need to note that \(N\) is locally contained in the graphs of such smooth functions, by the implicit function theorem. Then since \(N\) is second countable, it can be covered by countably many open sets each of which are contained in the graph of such a smooth function. ◻
Proposition 16. Let \(X\) be a compactum with \(\dim X \leq 1\), and let \(N \subset \mathbb{C}^n\) be a smooth submanifold of \(\mathbb{C}^n\) with real codimension at least \(2\). Then \(\mathcal{A}(X,N)\) is dense in \(C(X,\mathbb{C}^n)\).
Proof. Let \(M \subset \mathbb{C}^n\) denote the complex hypersurface \(z_1 z_2 \cdots z_n = 0\). By Lemma 15, we can find a sequence of autodiffeomorphisms \((f_k)_{k=1}^{\infty}\) of \(\mathbb{C}^n\) such that \(N \subset \bigcup_{k=1}^{\infty} f_k(\{z \in \mathbb{C}^n : z_n = 0\})\); hence \(N \subset \bigcup_{k=1}^{\infty} f_k(M)\). In particular, \[\begin{align} \bigcap_{k=1}^{\infty} \mathcal{A}(X,f_k(M)) \subset \mathcal{A}\left(X, \bigcup_{k=1}^{\infty} f_k(M) \right) \subset \mathcal{A}(X,N) \end{align}\] By Lemma 14, each \(\mathcal{A}(X,f_k(M))\) is an open dense subset of the Banach space \(C(X,\mathbb{C}^n)\), so by the Baire Category theorem, \(\bigcap_{k=1}^{\infty} \mathcal{A}(X,f_k(M))\) is a dense subset of \(C(X,\mathbb{C}^n)\). Hence by above, \(\mathcal{A}(X,N)\) is a dense subset of \(C(X,\mathbb{C}^n)\), as desired. ◻
We now apply the above to the discriminant variety \(V(\Delta)\), by decomposing it into a union of smooth manifolds, each with real codimension at least 2.
Corollary 17. Let \(X\) be a compactum with \(\dim X \leq 1\), and let \(V(\Delta) \subset \mathbb{C}^n\) denote the monic discriminant variety, i.e. \[V(\Delta)= \{ (a_0, a_1, \ldots, a_{n-1}) \in \mathbb{C}^n : \Delta(a_0 + a_1z + \cdots + a_{n-1}z^{n-1} + z^n) = 0 \}\] Then \(\mathcal{A}(X,V(\Delta))\) is an open dense subset of \(C(X,\mathbb{C}^n)\).
Proof. First, observe that \(V(\Delta)\) is the zero-set of a complex polynomial in \(n\) variables, and hence it has real codimension \(2\) and is a closed subset of \(\mathbb{C}^n\). By [14], the variety \(V(\Delta)\) admits a Whitney stratification \(D = D_2 \cup D_3 \cdots \cup D_n\) where each \(D_j\) consists of those points \((a_0, \ldots, a_{n-1}) \in V(\Delta)\) where the highest multiplicity root of the corresponding polynomial \(a_0 + a_1 z + \cdots + a_{n-1}z^{n-1} + z^n\) is \(j\). Each \(D_j\) is a real smooth submanifold of \(\mathbb{C}^n\) which is a closed subset \(V(\Delta)\) and has codimension at least that of \(V(\Delta)\); hence each has real codimension at least \(2\). It follows by Proposition 16 that \(\mathcal{A}(X, D_j)\) is a dense subset of \(C(X,\mathbb{C}^n)\) for each \(j\), and by Lemma 13, \(\mathcal{A}(X,D_j)\) is also open in \(C(X,\mathbb{C}^n)\). Since the intersection of finitely many open dense sets is open and dense, it follows that \(\mathcal{A}(X,V(\Delta)) = \bigcap_{j=2}^{n} \mathcal{A}(X,D_j)\) is open and dense in \(C(X,\mathbb{C}^n)\), as desired. ◻
From the above, we immediately obtain our perturbation theorem for one-dimensional continua.
Theorem 18. Let \(X\) be a compactum with \(\dim X \leq 1\), and let \(P:X \to \mathbb{C}^n\) be a monic polynomial over \(C(X)\) with degree \(n\). Then for all \(\varepsilon> 0\), there exists a monic polynomial over \(C(X)\) of degree \(n\) with no repeated roots \(Q: X \to B_n\) such that \(\left\Vert P-Q \right\Vert < \varepsilon\).
Proof. We need to show that \(C(X, B_n)\) is dense in \(C(X, \mathbb{C}^n)\), but the space \(C(X, B_n)\) is exactly \(\mathcal{A}(X, V(\Delta))\), so this follows immediately from Corollary 17 above. ◻
If we want to use this perturbation result to say something about approximate roots, we should also check that by perturbing the coefficients of a polynomial over \(C(X)\) does not change its approximate roots.
Lemma 19. Let \(X\) be a compactum, and let \(P(x,z) = z^n + a_{n-1}(x)z^{n-1} + \cdots + a_1(x)z + a_0(x)\) be a monic polynomial with coefficients in \(C(X)\). Then there exists a constant \(C > 0\) such that if \(\varepsilon\in (0,1]\) and if \(Q(x,z) = z^n + b_{n-1}(x)z^{n-1} + \cdots + b_1(x)z + b_0(x)\) is a monic polynomial with coefficients in \(C(X)\) such that \(\left\Vert a_k - b_k \right\Vert_{\infty} < \varepsilon\), then every \(\varepsilon\)-approximate solution of \(Q\) is a \(C\varepsilon\)-approximate solution of \(P\).
Proof. Let \(M = \max\{ \left\Vert a_0 \right\Vert, \left\Vert a_1 \right\Vert, \ldots, \left\Vert a_{n-1} \right\Vert \}\) and let \(C = \frac{(2+M)^n - 1}{1+M} + 1\). Take a monic polynomial \(Q(x,z) = z^n + b_{n-1}(x)z^{n-1} + \cdots + b_1(x)z + b_0(x)\) with coefficients in \(C(X)\) such that \(\left\Vert a_k - b_k \right\Vert_{\infty} < \varepsilon\). Let \(f \in C(X)\) be an \(\varepsilon\)-approximate solution of \(Q\). Then by Lemma 6, \(\left\Vert f \right\Vert \leq 1 + \varepsilon+ M \leq 2 + M\), so for each \(x \in X\), we have that \[\begin{align} |P(x,f(x))| &\leq |P(x,f(x)) - Q(x,f(x))| + |Q(x,f(x))| \\ &=|(a_{n-1}(x) - b_{n-1}(x))f(x)^{n-1} + \cdots + (a_1(x)-b_1(x))f(x) + (a_0(x)- b_0(x))| \\ & \qquad + |Q(x,f(x))| \\ &\leq \left\Vert a_{n-1} - b_{n-1} \right\Vert\left\Vert f \right\Vert^{n-1} + \cdots + \left\Vert a_1 - b_1 \right\Vert\left\Vert f \right\Vert + \left\Vert a_0 - b_0 \right\Vert + \varepsilon\\ &< \varepsilon(\left\Vert f \right\Vert^{n-1} + \cdots + \left\Vert f \right\Vert + 1) + \varepsilon\\ &\leq \varepsilon\left( (2+M)^{n-1} + \cdots + (2+M) + 1 \right) + \varepsilon\\ &= \varepsilon\left( \frac{(2+M)^n - 1}{(2+M) - 1} + 1 \right) \\ &= C\varepsilon \end{align}\] Since the above holds for all \(x \in X\), it follows that \(\sup_{x \in X} |P(x,f(x))| < C\varepsilon\), as desired. ◻
We are now ready to prove the main result of this section, Theorem 5.
Proof of Theorem 5. Clearly (1) implies (2), and (2) implies (3) by Theorem 10. Now assume that (3) holds, and take any monic polynomial \(P\) of degree at most \(n\) with coefficients in \(C(X)\). By Lemma 19, we can find a constant \(C > 0\) such that if \(\varepsilon_0 \in (0,1)\) and if \(Q\) is a monic polynomial with coefficients in \(C(X)\) such that \(\left\Vert P - Q \right\Vert < \varepsilon_0\), then every \(\varepsilon_0\)-approximate solution of \(Q\) is a \(C\varepsilon_0\)-approximate solution of \(P\). Now suppose that \(\varepsilon> 0\), and let \(\varepsilon_0 := \min(1, \frac{\varepsilon}{C})\). Since \(\dim X \leq 1\), by Theorem 18, we can find a monic polynomial \(Q\) with coefficients in \(C(X)\) having no repeated roots such that \(\deg Q = \deg P \leq n\) and \(\left\Vert P - Q \right\Vert < \varepsilon_0\). By assumption, \(Q\) has an exact root, say \(f \in C(X)\). Then in particular \(f\) is an \(\varepsilon_0\)-approximate root of \(Q\), so it follows by our choice of \(C\) that \(f\) is a \(C\varepsilon_0\)-approximate solution of \(P\). But \(\varepsilon\geq C \varepsilon_0\), so it follows that \(f\) is an \(\varepsilon\)-approximate solution of \(P\). This proves that (1) holds, as desired. ◻
Now Lemma 2.4 in [15] and the discussion thereafter show that if every monic polynomial over \(C(X)\) of degree at most \(n\) with no repeated roots has an exact root, then any monic polynomial \(P\) over \(C(X)\) with degree at most \(n\) can be factored completely, that is, we can find continuous functions \(\rho_1, \ldots, \rho_n \in C(X)\) such that \(P(x,z) = \prod_{j=1}^{n}(z-\rho_j(x))\). Thus as a consequence of Theorem 5 we obtain:
Corollary 20. Let \(X\) be a compactum with \(\dim X \leq 1\) and let \(n\) be a positive integer. The following are equivalent:
Every monic polynomial of degree at most \(n\) with coefficients in \(C(X)\) has approximate roots;
Every monic polynomial of degree at most \(n\) with coefficients in \(C(X)\) with no repeated roots has approximate roots;
Every monic polynomial of degree at most \(n\) with coefficients in \(C(X)\) with no repeated roots can be factored completely.
In the language of [5], the last condition in the corollary above says that the class of equations \(\overline{\mathfrak{A}_n}(X)\) consisting of monic polynomials over \(C(X)\) of degree at most \(n\) with no repeated roots is completely solvable.
To generalize the global arguments of homotopy theory to spaces that do not have nice local properties (such as local path-connectedness), we need shape theory. In this section, following [16], we recall some elements of shape theory and the language in which it is formulated, namely that of pro-categories.
Definition 21. Given a category \(\mathcal{C}\), we define its corresponding pro-category \(\operatorname{Pro}(\mathcal{C})\) to be the category consisting of the following objects and morphisms:
(i) The objects of \(\operatorname{Pro}(\mathcal{C})\) are given by inverse systems \((X_\alpha, p_{\alpha}^{\alpha'}, A)\), which are functors from a directed set \(A\) to \(\mathcal{C}\). Explicitly, for each \(\alpha \in A\) we have an object \(X_\alpha\) of \(\mathcal{C}\), and for each pair \(\alpha' \leq \alpha\) we have a morphism \(p_{\alpha}^{\alpha'}: X_{\alpha'} \to X_\alpha\).
(ii) The set of morphisms in \(\operatorname{Pro}(\mathcal{C})\) between two objects \(\underline{X} = (X_\alpha, p_{\alpha}^{\alpha'}, A)\) and \(\underline{Y} = (Y_\beta, q_{\beta}^{\beta'}, B)\) is given by \[\operatorname{Hom}_{\operatorname{Pro}(\mathcal{C})}(\underline{X}, \underline{Y}) = \varprojlim_{\beta \in B} \varinjlim_{\alpha \in A} \operatorname{Hom}_{\mathcal{C}}(X_\alpha, Y_\beta)\]
It is worth noting a few remarks concerning the above definition. To start, notice that we have a fully faithful inclusion \(\mathcal{C}\to \operatorname{Pro}(\mathcal{C})\) by using indexing sets of a single element: \(X \mapsto (X, \operatorname{id}_X, *)\). Under this functor, the image of the inverse system \((X_\alpha, p_{\alpha}^{\alpha'}, A)\) in \(\mathcal{C}\) is an inverse system in \(\operatorname{Pro}(\mathcal{C})\), whose projective limit in \(\operatorname{Pro}(\mathcal{C})\) is precisely the object \(\underline{X} = (X_\alpha, p_{\alpha}^{\alpha'}, A)\). This perspective lets us understand morphisms into an object \(\underline{Y} = (Y_\beta, q_{\beta}^{\beta'}, B)\) of \(\operatorname{Pro}(\mathcal{C})\) by understanding a system of morphisms into objects \(Y_\beta\) of \(\mathcal{C}\): \[\begin{align} \operatorname{Hom}_{\operatorname{Pro}(\mathcal{C})}(\underline{X}, \underline{Y}) & = \operatorname{Hom}_{\operatorname{Pro}(\mathcal{C})}(\underline{X}, \varprojlim_{\beta \in B} Y_\beta) \\ & = \varprojlim_{\beta \in B} \operatorname{Hom}_{\operatorname{Pro}(\mathcal{C})}(\underline{X}, Y_\beta) \end{align}\]
Unlike for the second entry, we cannot in general pull the projective limit out of the first entry. Thus the following identity \[\operatorname{Hom}_{\operatorname{Pro}(\mathcal{C})}(\underline{X}, Y) = \varinjlim_{\alpha \in A} \operatorname{Hom}_{\mathcal{C}}(X_\alpha, Y)\] is really a definition. Note that an element \(f\) of this direct limit is an equivalence class of morphisms, with representatives given by morphisms \(f_\alpha: X_\alpha \to Y\) for some \(\alpha \in A\); two representatives \(f_\alpha \in \operatorname{Hom}_{\mathcal{C}}(X_\alpha, Y)\) and \(f_{\alpha'} \in \operatorname{Hom}_{\mathcal{C}}(X_{\alpha'}, Y)\) represent the same class if there exists some \(\alpha'' \in A\) with \(\alpha \leq \alpha''\), and \(\alpha' \leq \alpha''\), and \[f_\alpha \circ p_{\alpha}^{\alpha''} = f_{\alpha'} \circ p_{\alpha'}^{\alpha''}\]
It is worth noting that the assignment \(\mathcal{C}\mapsto \operatorname{Pro}(\mathcal{C})\) is a functor, where a functor \(\mathcal{F}: \mathcal{C}\to \mathcal{D}\) is sent to the functor \(\operatorname{Pro}(\mathcal{F})\) that applies \(\mathcal{F}\) to all objects and structure maps of the inverse system. In particular, a subcategory \(\mathcal{D}\subseteq\mathcal{C}\) gives rise to a subcategory \(\operatorname{Pro}(\mathcal{D}) \subseteq\operatorname{Pro}(\mathcal{C})\).
The goal of shape theory is to understand homotopy classes of maps from a general topological space into a CW complex. Let us denote the category of topological spaces as \(\mathsf{Spc}\) and the full subcategory of CW complexes by \(\mathsf{CW}\). We shall denote their homotopy categories by \(\operatorname{Ho}\mathsf{Spc}\) and \(\operatorname{Ho}\mathsf{CW}\) respectively.
Definition 22. Let \(X\) be a topological space. Suppose we have an object \(\underline{W} = (W_\alpha, p_{\alpha}^{\alpha'}, A) \in \operatorname{Pro}(\operatorname{Ho}\mathsf{CW})\) along with a map \(\underline{q}: X \to \underline{W}\) in \(\operatorname{Pro}(\operatorname{Ho}\mathsf{Spc})\), which is just a collection of maps \(q_\alpha: X \to W_\alpha\) respecting the structure maps \(p_{\alpha}^{\alpha'}\). We say that (\(\underline{W}\), \(\underline{q}\)) is the shape of \(X\) if \(\underline{q}\) is initial among maps from \(X\) to objects of \(\operatorname{Pro}(\operatorname{Ho}\mathsf{CW})\), meaning for any other \(\underline{Z} \in \operatorname{Pro}(\operatorname{Ho}\mathsf{CW})\) and \(\underline{f}: X \to \underline{Z}\) there is a unique morphism \(\underline{g}: \underline{W} \to \underline{Z}\) making the diagram commute: \[\begin{tikzcd} X \\ {\underline{W}} & {\underline{Z}} \arrow["{\underline{q}}"', from=1-1, to=2-1] \arrow["{\underline{f}}", from=1-1, to=2-2] \arrow["{\underline{g}}"', dashed, from=2-1, to=2-2] \end{tikzcd}\] This universal property makes the shape of \(X\) well-defined up to natural isomorphism in \(\operatorname{Pro}(\operatorname{Ho}\mathsf{CW})\).
Remark 23. Equivalently, (\(\underline{W}\), \(\underline{q}\)) is the shape of \(X\) if and only if any morphism \(f: X \to K\) with \(K\) a CW complex factors uniquely through \(\underline{W}\) in \(\operatorname{Pro}(\operatorname{Ho}\mathsf{Spc})\). Indeed, since \(\operatorname{Ho}\mathsf{CW}\subseteq\operatorname{Pro}(\operatorname{Ho}\mathsf{CW})\) this is implied by the original definition, and conversely given a factorization along single CW complexes and \(\underline{f}: X \to \underline{Z} \in \operatorname{Pro}(\operatorname{Ho}\mathsf{CW})\), for each \(f_\beta: X \to Z_\beta\) we get a \(g_\beta: \underline{W} \to Z_\beta\) making the diagram commute \[\begin{tikzcd} X \\ {\underline{W}} & {Z_\beta} \arrow["{\underline{q}}"', from=1-1, to=2-1] \arrow["{f_\beta}", from=1-1, to=2-2] \arrow["{g_\beta}"', dashed, from=2-1, to=2-2] \end{tikzcd}\] which by uniqueness give that the maps \(g_\beta \in \operatorname{Hom}(\underline{W}, Z_\beta)\) assemble to a map \(\underline{g} \in \operatorname{Hom}(\underline{W}, \underline{Z})\) with \(\underline{f} = \underline{g} \circ \underline{q}\) as in the original definition.
Remark 24. We can instead work with the pointed homotopy category \(\operatorname{Ho}\mathsf{Spc}_*\) and the subcategory \(\operatorname{Ho}\mathsf{CW}_*\) of pointed CW complexes with basepoint-preserving maps up to basepoint-preserving homotopy. In this case, the pointed shape \((\underline{W}, w_0) \in \operatorname{Pro}(\operatorname{Ho}\mathsf{CW}_*)\) of a pointed space \((X, x_0)\) is given by the same definition as above, by appropriately making all objects and morphisms pointed.
As seen in Remark 23, the shape of \(X\) captures all of the data about mapping out of \(X\) and into CW complexes up to homotopy. This can be made precise by considering open coverings of \(X\) and partitions of unity as in [16]. In particular, as a consequence of Theorem 3.1.4 in [16] we obtain:
Lemma 25. Given a topological space \(X\), there is a \(\underline{W} \in \operatorname{Pro}(\operatorname{Ho}\mathsf{CW})\) and a morphism \(\underline{q}: X \to \underline{W}\) of \(\operatorname{Pro}(\operatorname{Ho}\mathsf{Spc})\) such that \((\underline{W}, \underline{q})\) is the shape of \(X\). Analogously, every pointed topological space has a pointed shape.
This lets us define the shape functor \(\operatorname{Shape}: \mathsf{Spc}\to \operatorname{Pro}(\operatorname{Ho}\mathsf{CW})\) taking a topological space to its shape (strictly speaking, we should pick a constructive way of assigning an inverse system of CW complexes to a space \(\mathsf{Spc}\to \operatorname{Pro}(\mathsf{CW})\), which represents the shape once we take the maps up to homotopy. One choice is to use Čech systems, as in [16]). In this paper we will require a way of computing the shape of a compactum, in order to perform calculations. This is provided to us via the following theorem (Theorem 4.1.5 in [16]):
Theorem 26. Consider an inverse system \(\underline{X} = (X_\alpha, p_{\alpha}^{\alpha'}, A)\) where each \(X_\alpha\) is a finite CW complex, and its limit \[X = \varprojlim_{\alpha \in A} X_\alpha\] Then \((\underline{X}, \underline{p})\) is the shape of \(X\), where \(\underline{p} = \{p_\alpha : X \to X_\alpha\}\) are the structure maps provided by the inverse limit. The analogous statement holds for inverse systems of pointed finite CW complexes.
We can combine this result with a theorem of Freudenthal [17] which says that we may write any metrizable compactum as an inverse limit of a sequence of polyhedra with piecewise-linear maps.
Finally we may use this framework to define some topological invariants.
Definition 27. Given any functor which is a homotopy invariant \(\mathcal{F}: \operatorname{Ho}\mathsf{CW}\to \mathsf{Grp}\), we can turn it into a shape invariant by defining: \[\begin{align} \underline{\mathcal{F}}: \mathsf{Spc}& \longrightarrow \operatorname{Pro}(\mathsf{Grp}) \\ X & \longmapsto \operatorname{Pro}(\mathcal{F})(\operatorname{Shape}(X)) \end{align}\] Note that the shape \(\operatorname{Shape}(X)\) lies in \(\operatorname{Pro}(\mathsf{CW})\), and then we may apply the pro-version of \(\mathcal{F}\), which is a functor \(\operatorname{Pro}(\mathcal{F}): \operatorname{Pro}(\operatorname{Ho}\mathsf{CW}) \to \operatorname{Pro}(\mathsf{Grp})\). Analogously, if we have a pointed homotopy invariant \(\mathcal{F}: \operatorname{Ho}\mathsf{CW}_* \to \mathsf{Grp}\) we can define the pointed shape invariant \(\underline{\mathcal{F}}: \mathsf{Spc}_* \to \operatorname{Pro}(\mathsf{Grp})\).
In this fashion, we obtain homotopy pro-groups and (co)homology pro-groups \(\underline{\pi}_n\), \(\underline{H}_n\), and \(\underline{H}^n\) by applying the above construction to the usual functors for homotopy groups \(\pi_n\), homology groups \(H_n\), and cohomology groups \(H^n\) (note that \(\underline{H}^n\) actually takes values in \(\operatorname{Ind}(\mathsf{Grp}) = \operatorname{Pro}(\mathsf{Grp}^{op})\)).
It is common to distill the invariants to get a group valued invariant. Note that \(\underline{H}_n(X)\) (resp. \(\underline{H}^n(X)\)) consist of a projective (resp. inductive) system of groups, and hence we may take an inverse (resp. direct) limit to get a single group. In this way, we obtain the Čech homology and cohomology groups: \[\check{H}_n(X) = \varprojlim \underline{H}_n(X) \qquad \check{H}^n(X) = \varinjlim \underline{H}^n(X)\]
In our investigation of continua we’ll see that the key invariant is the fundamental pro-group \(\underline{\pi}_1(X, x_0)\), and we will mostly use the other invariants as a way to gain information about \(\underline{\pi}_1(X, x_0)\) or answer more classical questions (usually posed in terms of \(\check{H}^1(X)\)).
Recall that given \(n \geq 1\) and a group \(G\), which has to be abelian if \(n \geq 2\), we can find an explicit pointed space \(K(G, n)\), called an Eilenberg-MacLane space, whose homotopy groups vanish except for \(\pi_n(K(G, n), x_0) = G\). Moreover this assignment taking in (discrete) groups and giving pointed CW complexes is functorial: \[\begin{align} {3} K(-, 1) & : \mathsf{Grp}&& \longrightarrow \mathsf{Spc}_* \\ K(-, n) & : \mathsf{Ab}&& \longrightarrow \mathsf{Spc}_* \qquad \text{for} \;n \geq 2 \end{align}\] This is nicely described in [18]. We can always take a model for \(K(G, n)\) as a pointed CW complex up to homotopy equivalence. We will mostly be interested in \(n = 1\).
One very useful property of Eilenberg-MacLane spaces is that we can completely characterize maps into them up to homotopy. We have the following correspondence:
Proposition 28. [19] Let \((X, x_0)\) be a pointed CW complex. Then the forgetful map \[\pi_1 : [(X, x_0), (K(G, 1), y_0)] \longrightarrow \operatorname{Hom}_{\mathsf{Grp}}(\pi_1(X, x_0), G)\] is a bijection. That is, pointed homotopy classes of maps into \(K(G, 1)\) are fully determined by the induced map on fundamental groups.
If we use a CW complex to model the Eilenberg-MacLane space, then we can upgrade this correspondence to continua.
Lemma 29. Let \((X, x_0)\) be a pointed continuum. If \((Y, y_0)\) is a CW model of an Eilenberg-MacLane space \(K(G, 1)\), then the forgetful map \[\underline{\pi}_1 : [(X, x_0), (Y, y_0)] \longrightarrow \operatorname{Hom}_{\operatorname{Pro}\mathsf{Grp}}(\underline{\pi}_1(X, x_0), G)\] is a bijection. That is, pointed homotopy classes of maps into \(Y\) are fully determined by the induced map on fundamental pro-groups.
Proof. Take an inverse system \(\underline{W} = (W_\alpha, w_\alpha, p_{\alpha}^{\alpha'}, A)\) of pointed connected CW complexes and a pointed morphism \(\underline{q}: (X, x_0) \to (\underline{W}, \underline{w})\) so that \((\underline{W}, \underline{q})\) is the pointed shape of \((X, x_0)\).
We start by proving surjectivity. Given a map \(\underline{\varphi}: \underline{\pi}_1(X, x_0) \to G\), we know it is represented by a group homomorphism \(\varphi_\beta : \pi_1(W_\beta, w_\beta) \to G\), for some index \(\beta \in A\). Now applying Proposition 28, this homomorphism is induced by a pointed map \(f_\beta: (W_\beta, w_\beta) \to (Y, y_0)\). But then \(f_\beta \circ q_\beta: (X, x_0) \to (Y, y_0)\) is a map such that \(\underline{\pi}_1(f_\beta \circ q_\beta) = \underline{\varphi}\).
Now for injectivity, consider two pointed maps \(f, g: (X, x_0) \to (Y, y_0)\) such that \(\underline{\pi}_1(f) = \underline{\pi}_1(g)\). Using the definition of the shape of \(X\), there is an index \(\beta \in A\) and pointed maps \(f_\beta, g_\beta: (W_\beta, w_\beta) \to (Y, y_0)\) such that \(f_\beta \circ q_\beta\) is base point homotopic to \(f\) and \(g_\beta \circ q_\beta\) is base point homotopic to \(g\). Since \(\underline{\pi}_1(f) = \underline{\pi}_1(g)\), there is a \(\gamma \geq \beta\) such that the map \(\pi_1(f_\beta \circ p_\beta^\gamma) : \pi_1(W_\gamma. w_\gamma) \to G\) is equal to the map \(\pi_1(g_\beta \circ p_\beta^\gamma)\). But since the pointed maps \(f_\gamma = f_\beta \circ p_\beta^\gamma\) and \(g_\gamma = g_\beta \circ p_\beta^\gamma\) from \((W_\gamma, w_\gamma)\) to \((Y, y_0) \cong K(G, 1)\) induce the same map on \(\pi_1\), they must be base point homotopic by Proposition 28. But given such a homotopy \(f_\gamma \simeq g_\gamma\), we obtain that \[\begin{align} f & \simeq f_\beta \circ q_\beta \simeq f_\beta \circ p_\beta^\gamma \circ q_\gamma = f_\gamma \circ q_\gamma \simeq g_\gamma \circ q_\gamma = g_\beta \circ p_\beta^\gamma \circ q_\gamma \simeq g_\beta \circ q_\beta \simeq g \end{align}\] and so \(f\) and \(g\) are base point homotopic. ◻
The functor \(K(-, 1): \mathsf{Grp}\rightarrow \mathsf{Spc}_*\) is a right inverse for the fundamental group \(\pi_1: \mathsf{Spc}_* \to \mathsf{Grp}\), and it would be useful to have something similar for the fundamental pro-group. We could start out by applying the functor \(\operatorname{Pro}\): \[\operatorname{Pro}(K(-, 1)): \operatorname{Pro}(\mathsf{Grp}) \longrightarrow \operatorname{Pro}(\mathsf{Spc}_*)\] but this results in a pro-space. However, as mentioned above, up to homotopy we can replace each \(K(G_\alpha, 1)\) with a CW model, which is one step closer as then we’ve constructed a shape with the desired property. Still the problem remains of finding a concrete space with this given shape. If we were somehow guaranteed that each \(K(G_\alpha, 1)\) were a finite CW complex, then Theorem 26 would give us a space by simply taking the inverse limit.
We seem to be asking for too much, as it is uncommon for \(K(G, 1)\) to be modeled by a finite CW complex. Indeed, if all we are asking for is an inverse to \(\underline{\pi}_1\), then we need not worry about higher homotopy groups and hence do not need Eilenberg-MacLane spaces. Given any finitely presented group \(G\) we can form a finite \(2\)-dimensional pointed CW complex \((X, x_0)\) with \(\pi_1(X, x_0) = G\) by using a \(1\)-cell for each generator and a \(2\)-cell for each relation. If \(G\) is a finitely generated free group, then \((X, x_0)\) can be taken to be \(1\)-dimensional (a wedge of finitely many circles).
Thus, given a pro-group \(\underline{G} = (G_\alpha, p_{\alpha}^{\alpha'}, A)\) for each \(G_\alpha\) we can find a finite \(2\)-dimensional pointed CW complex \((X_\alpha, x_\alpha)\) whose fundamental group is \(G_\alpha\), and for each \(p_{\alpha}^{\alpha'}: G_{\alpha'} \to G_\alpha\) we can choose a pointed (cellular) map \(\varphi_{\alpha}^{\alpha'}: (X_{\alpha'}, x_{\alpha'}) \to (X_\alpha, x_\alpha)\) that induces \(p_{\alpha}^{\alpha'}\) on fundamental groups (though this choice is not unique up to homotopy). Then by Theorem 26, the inverse limit \[(X, x_0) = \varprojlim_{\alpha \in A} (X_\alpha, x_\alpha)\] is a \(2\)-dimensional pointed continuum with \(\underline{\pi}_1(X, x) = \underline{G}\). We summarize the above discussion in the following lemma.
Lemma 30. Given a pro-group \(\underline{G}\) lying in the subcategory \(\operatorname{Pro}(\mathsf{Grp}^{\mathrm{fp}})\) of inverse systems of finitely presented groups, there exists a \(2\)-dimensional pointed continuum \((X, x_0)\) with the fundamental pro-group \(\underline{\pi}_1(X, x_0) = \underline{G}\). Moreover, we may take \((X, x_0)\) to be \(1\)-dimensional if \(\underline{G}\) lies in the subcategory \(\operatorname{Pro}(\mathsf{Grp}^{\mathrm{fg}, \mathrm{free}})\) of inverse systems of finitely generated free groups.
As seen in Section 2.2, a polynomial with no repeated roots on \(X\) is nothing but a map \(P: X \to B_n\), and solving these polynomials amounts to lifting across the covering map \(E_n \to B_n\). Recall also that the spaces \(B_n\) and \(E_n\) happen to be Eilenberg-MacLane spaces for the groups \(\mathcal{B}_n\) and \(M_n\) respectively, so we already know a lot about maps into these spaces.
Corollary 31. Consider a pointed continuum \((X, x_0)\) and a degree \(n\) monic polynomial \(b_0 \in B_n\) and a root \(\lambda_0 \in E_n\) of \(b_0\).
(i) For each map of pro-groups \(\underline{\varphi} : \underline{\pi}_1(X, x_0) \to \mathcal{B}_n\), there exists a polynomial \(P: (X, x_0) \to (B_n, b_0)\) with \(\underline{\pi}_1(P) = \underline{\varphi}\).
(ii) Consider a polynomial \(P: (X, x_0) \to (B_n, b_0)\), with its \(n\) distinct roots at \(x_0\) given as \(\{z_1, \dots, z_n\}\). Then \(P\) has a solution \(\lambda: (X, x_0) \to (E_n, z_i)\) if and only if \(\underline{\pi}_1(P)\) has its image lying in the subgroup \(\pi_1(E_n, z_i) \cong M_{n,i}\) of \(\mathcal{B}_n \cong \pi_1(B_n, b_0)\). Moreover, \(P\) is completely solvable if and only if \(\underline{\pi}_1(P)\) has its image lying in the subgroup \(N_n\) of \(\mathcal{B}_n \cong \pi_1(B_n, b_0)\).
Proof. All are a consequence of Lemma 29. Note that \(B_n\) and \(E_n\) are both smooth manifolds and hence they are CW models for the Eilenberg-MacLane spaces \(K(\mathcal{B}_n, 1)\) and \(K(M_n, 1)\) respectively.
To prove (i), we use that the forgetful map \[\underline{\pi}_1 : [(X, x_0), (B, b_0)] \longrightarrow \operatorname{Hom}_{\operatorname{Pro}\mathsf{Grp}}(\underline{\pi}_1(X, x_0), \mathcal{B}_n)\] is surjective to find a polynomial \(P\) that maps to \(\underline{\varphi}\).
For (ii), the “if" direction is immediate, as \(P\) having a solution \(\lambda\) means that the diagram \[\begin{tikzcd} & (E_n, z_i) \\ (X, x_0) & (B_n, b_0) \arrow["P"', from=2-1, to=2-2] \arrow[from=1-2, to=2-2] \arrow["\lambda", from=2-1, to=1-2] \end{tikzcd}\] commutes, which, after applying the functor \(\underline{\pi}_1\), gives the commutative diagram \[\begin{tikzcd} & M_{n,i} \\ \underline{\pi}_1(X, x_0) & \mathcal{B}_n \arrow["\underline{\pi}_1(P)"', from=2-1, to=2-2] \arrow[hook, from=1-2, to=2-2] \arrow["\underline{\pi}_1(\lambda)", from=2-1, to=1-2] \end{tikzcd}\] showing that the image of \(\underline{\pi}_1(P)\) lies in \(M_{n,i}\).
For the converse, assume that the image of \(\underline{\pi}_1(P)\) lies in \(M_{n,i}\), which means that we have a commutative diagram \[\begin{tikzcd} & M_{n,i} \\ \underline{\pi}_1(X, x_0) & \mathcal{B}_n \arrow["\underline{\pi}_1(P)"', from=2-1, to=2-2] \arrow[hook, from=1-2, to=2-2] \arrow["\underline{\varphi}", from=2-1, to=1-2] \end{tikzcd}\] where \(\underline{\varphi}: \underline{\pi}_1(X, x_0) \to M_{n,i}\) is the map given by shrinking the codomain of \(\underline{\pi}_1(P)\). Similarly to part (i), we can use surjectivity of the map in Lemma 29 for the Eilenberg-MacLane space \(E_n = K(M_{n,i}, 1)\) to get a pointed map \(f: (X, x_0) \to (E_n, z_i)\) with \(\underline{\pi}_1(f) = \underline{\varphi}\).
Letting \(\rho: (E_n, z_i) \to (B_n, b_0)\) denote the covering map, we see that \(\rho \circ f: (X, x_0) \to (B_n, b_0)\) is a pointed map such that \[\underline{\pi}_1(\rho \circ f) = \underline{\pi}_1(P)\] Using injectivity of the map in Lemma 29 for the Eilenberg-MacLane space \(B_n\), we see that \(\rho \circ f\) is homotopic to \(P\) (relative to the basepoint). Finally, since \(\rho\) is a covering map, the property of having a lift across \(\rho\) is a homotopy invariant (by lifting the homotopy), so \(\rho \circ f\) having a lift \(f\) implies that \(P\) has a lift \(\lambda\) as required.
Now we consider the statement about complete solvability. The polynomial \(P\) being completely solvable is equivalent to there being \(n\) solutions \(\lambda_i\) on \(X\), so that their values at \(x_0\) are the \(n\) distinct complex numbers \(z_i\); that is, there exist solutions \(\lambda_i: (X, x_0) \to (E_n, z_i)\) for all \(1 \leq i \leq n\). By the first part of (ii) we know that this is equivalent to \(\underline{\pi}_1(P)\) having its image in all of the \(M_{n, i}\). Since the intersection of the \(M_{n, i}\) is \(N_n\) this completes the proof. ◻
This Corollary inspires the following conditions a pro-group \(\underline{G}\) can satisfy:
For any map \(\underline{\varphi}: \underline{G} \to B_n\), the image of \(\underline{\varphi}\) lies in one of the \(M_{n,i}\).
For any map \(\underline{\varphi}: \underline{G} \to B_n\), the image of \(\underline{\varphi}\) lies in \(N_n = \ker(\tau)\), where \(\tau : \mathcal{B}_n \to S_n\) is the canonical map. Equivalently, for any map \(\underline{\varphi}: \underline{G} \to B_n\) the composition \(\tau \circ \underline{\varphi}: \underline{G} \to S_n\) is trivial.
We now have complete topological characterizations of when monic polynomials with no repeated roots over \(C(X)\) have exact solutions. Along with our results in section 3, for the case \(\dim X \leq 1\) we also have characterizations of when monic polynomials \(C(X)\) have approximate roots.
Theorem 32. Let \(X\) be a continuum and \(n\) be a positive integer. Then
(i) Every monic polynomial of degree \(n\) over \(C(X)\) with no repeated roots has an exact root if and only if \(\underline{\pi}_1(X)\) satisfies \((*_n)\).
(ii) Every monic polynomial of degree \(n\) over \(C(X)\) with no repeated roots can be factored completely if and only if \(\underline{\pi}_1(X)\) satisfies \((**_n)\).
(iii) Given a pro-group \(\underline{G}\) which is an inverse system of finitely presented groups, and a morphism \(\underline{\varphi}: \underline{G} \to B_n\) whose image does not lie in any of the \(M_{n, i}\), there exists a \(2\)-dimensional pointed continuum \((X, x_0)\) with \(\underline{\pi}_1(X, x_0) \cong G\) and a polynomial \(P:X \to B_n\) which has no exact root (and hence no approximate roots in the case \(\dim X \leq 1\)).
Applying Theorem 32 and Corollary 20 for each positive integer, we obtain a topological characterization for when the ring of continuous functions \(C(X)\) is approximately algebraically closed in terms of the fundamental pro-group.
Theorem 33. Let \(X\) be a continuum with \(\dim X \leq 1\). The following are equivalent:
(i) \(C(X)\) is approximately algebraically closed.
(ii) The fundamental pro-group \(\underline{\pi}_1(X)\) satisfies \((*_n)\) for every \(n\).
(iii) The fundamental pro-group \(\underline{\pi}_1(X)\) satisfies \((**_n)\) for every \(n\).
Example 1. The notion of a co-existentially closed continuum, first introduced by Bankston in [9], is important in model theory. As shown in [10], co-existentially closed continua are one-dimensional and hereditarily indecomposable. It has recently been shown that co-existentially closed continua are approximately algebraically closed [8]. Thus, Theorem 33 applies, and we obtain a new topological property of co-existentially closed continua, concerning their fundamental pro-group.
We now show that in order to solve certain special polynomials, namely those of the form \(z^m - f \in C(X)[z]\) for \(f\) non-vanishing, it is enough to look at the Čech cohomology of \(X\). For this, we will need to recall the following property of groups.
Definition 34. A group \(G\) is called \(m\)-divisible if for every \(g \in G\) there exists some \(h \in G\) such that \(h^m = g\). We shall call \(G\) divisible if it is \(m\)-divisible for every integer \(m \geq 1\).
In particular, there are some classical results are stated in terms of the divisibility of \(\check{H}^1(X)\), for which we now provide new, updated proofs using shape-theoretic invariants. We start with a small proposition.
Proposition 35. Given a pro-group \(\underline{G}\), consider the dual group \[A = \varinjlim \operatorname{Hom}(\underline{G}, \mathbb{Z})\] Then \(A\) is \(m\)-divisible if and only if any morphism \(\underline{\varphi}: \underline{G} \to \mathbb{Z}\) has its image land in the subgroup \(m\mathbb{Z}\) (and hence \(m^k\mathbb{Z}^d\) for each \(k\), by induction).
Proof. A morphism \(\underline{\varphi}: \underline{G} \to \mathbb{Z}\) is exactly an element of \(A\), and its image lies in \(m\mathbb{Z}\) if and only if there exists some index \(\alpha\) such that the image of a representative \(\varphi_\alpha : G_\alpha \to \mathbb{Z}\) lands in \(m\mathbb{Z}\), if and only if there exists some index \(\alpha\) and homomorphism \(\psi_\alpha : G_\alpha \to \mathbb{Z}\) such that \(\varphi_\alpha = m\psi_\alpha\), if and only if there exists some \(\underline{\psi}: \underline{G} \to \mathbb{Z}\) with \(m \underline{\psi} = \underline{\varphi}\), which is the definition of \(A\) being \(m\)-divisible. ◻
With this we can analyze \(m\)th roots of non-vanishing functions, and we arrive at a well-known statement in the spirit of [2].
Corollary 36. Given a continuum \(X\), we get that \(\check{H}^1(X)\) is \(m\)-divisible if and only if for any non-vanishing \(f \in C(X)\) there exists a \(g \in C(X)\) such that \(g^m = f\).
Proof. Consider the subset \(C_m \subseteq B_m\) consisting of polynomials \(z^m - \mu\) for \(\mu \in \mathbb{C}^\times\), which by this parametrization is homeomorphic to \(\mathbb{C}^\times\). Then the preimage of \(C_m\) under our usual covering map \(E_m \to B_m\) is again a copy of \(\mathbb{C}^\times\), and the covering map is given by \(\lambda \mapsto \lambda^m\). Then given such an \(f\) we get a polynomial \(P_f: X \to C_m \subseteq B_m\) by \(P_f(x, z) = z^m - f(x)\) and the question of asking for a solution \(g\) is equivalent to asking for a lift on fundamental groups: \[\begin{tikzcd} & {\mathbb{C}^\times \subseteq E_m} &&& {m\mathbb{Z}} \\ X & {C_m \subseteq B_m} && {\underline{\pi}_1(X,x_0)} & \mathbb{Z} \arrow[from=1-2, to=2-2] \arrow[hook, from=1-5, to=2-5] \arrow[dashed, from=2-1, to=1-2] \arrow["{P_f}"', from=2-1, to=2-2] \arrow[dashed, from=2-4, to=1-5] \arrow["{\underline{\pi}_1(P_f)}"', from=2-4, to=2-5] \end{tikzcd}\] Now, by Proposition 35 since the dual group \[\check{H}^1(X) = \varinjlim \operatorname{Hom}(\underline{\pi}_1(X,x_0), \mathbb{Z})\] is \(m\)-divisible, we get that the image of \(\underline{\pi}_1(P_f)\) lies in \(m\mathbb{Z}\) as required.
Conversely, given a morphism \(\underline{\varphi}: \underline{\pi}_1(X,x_0) \to \mathbb{Z}\) we know it is represented by some \(\varphi_\alpha: \pi_1(W_\alpha, w_\alpha) \to \mathbb{Z}\), where \(\underline{q}: (X, x_0) \to (\underline{W}, \underline{w})\) is the pointed shape of \((X,x_0)\). But \(C_m \cong \mathbb{C}^\times\) is an Eilenberg-MacLane space \(K(\mathbb{Z}, 1)\), so this \(\varphi_\alpha\) is represented by a pointed map \(h_\alpha: (W_\alpha, w_\alpha) \to (\mathbb{C}^\times, 1)\) and \(P_f = h_\alpha \circ q_\alpha: X \to C_m\) is a polynomial \(P_f(z, x) = z^m - f(x)\) with \(\underline{\pi}_1(P_f) = \underline{\varphi}\). By assumption \(f\) has an \(m^{\text{th}}\) root \(g\), which by the above reasoning means that \(\underline{\pi}_1(P_f) = \underline{\varphi}\) lands in \(m\mathbb{Z}\). So \(\check{H}^1(X)\) is \(m\)-divisible by Proposition 35. ◻
The next theorems are about using \(m^{\text{th}}\) roots to solve more complicated polynomials. All of these results are about assuming some amount of solvability of \(\underline{\pi}_1(X)\). The first is a restatement of [5] in our language of fundamental pro-groups, dealing with the case that the fundamental pro-group is an inverse limit of abelian groups.
Theorem 37. Let \(X\) be a continuum with \(\dim X \leq 1\) such that the fundamental pro-group \(\underline{\pi}_1(X)\) is an inverse limit of abelian groups. The following are equivalent:
(i) \(C(X)\) is approximately algebraically closed.
(ii) The first Čech cohomology group \(\check{H}^1(X)\) is divisible.
Proof. By Corollary 20, (i) is equivalent to the assertion that every monic polynomial with coefficients in \(C(X)\) with no repeated roots can be factored completely. But by [5], since \(\underline{\pi}_1(X)\) is an inverse limit of abelian groups, this is equivalent to (ii). ◻
Examples 38. Using Theorem 37 above, we can describe some examples and non-examples of approximately algebraically closed continua.
A tree-like continuum is approximately algebraically closed.
A solenoid \(\Sigma\) is approximately algebraically closed if and only if it is the universal one (i.e. has \(\check{H}^1(\Sigma; \mathbb{Z}) = \mathbb{Q}\)).
A pseudo-solenoid \(\mathbb{P}\Sigma\) is approximately algebraically closed if and only if it is the universal one.
Descriptions and definitions of solenoids and pseudo-solenoids may be found in [8]. It was observed in [15] that a solenoid or pseudo-solenoid which is not universal cannot be approximately algebraically closed; Theorem 37 establishes the converse of this observation, and gives further indication that the universal pseudo-solenoid may be co-existentially closed, which is partial progress towards answering [8].
Example 2. There exists a one-dimensional continuum \(X\) which is acyclic (i.e. \(\check{H}^1(X) = 0\)), but for which the fundamental pro-group \(\underline{\pi}_1(X)\) is non-abelian (i.e. cannot be written as an inverse limit of abelian groups) and satisfies \((*_n)\) for every \(n \geq 1\). In particular \(C(X)\) is approximately algebraically closed, but we cannot use the criterion in Theorem 37 to determine this, and must use the full power of the main theorem.
To produce such an \(X\), we will first construct a sequence of nested subgroups \[\begin{align} G_1 \supseteq G_2 \supseteq G_3 \supseteq \cdots \end{align}\] with the following properties:
\(G_n\) is a free group on two generators for every \(n\);
\(G_{n+1}\) is contained in the commutator subgroup \(G_n'\) of \(G_{n}\) for each \(n\); and,
Given any group homomorphism \(\varphi : G_n \to \mathcal{B}_k\) from any group \(G_n\) in the sequence to any braid group \(\mathcal{B}_k\), there is some integer \(m \geq n\) such that \(G_m\) lies in the kernel of \(\tau \circ \varphi\), where \(\tau : \mathcal{B}_k \to S_k\) is the canonical map.
We will construct these groups recursively, and at each stage of the recursion, we will not only produce a group \(G_n\), but we will also keep track of an enumeration of the countable set \(\bigcup_{k=1}^{\infty}\operatorname{Hom}(G_n,\mathcal{B}_k)\). This enumeration will help build the groups further down in the sequence. We will also let \((\ell_i)_{i =1}^{\infty}\) be a sequence of positive integers with the property that each positive integer appears infinitely often in the sequence and \(\ell_i \leq i\) for all \(i\). For example, we could take \((\ell_i)\) to be the sequence \(1,1,2,1,2,3,1,2,3,4, \ldots\).
For the initial step in the recursion, set \(G_1 = \mathbb{Z} * \mathbb{Z}\), and choose an enumeration \((\varphi_{1,j})_{j=1}^{\infty}\) of the set \(\bigcup_{k=1}^{\infty} \operatorname{Hom}(G_1, \mathcal{B}_k)\). Note that this set is actually countable because \(\mathcal{B}_k\) is countable and \(G_1\) is finitely generated. For the recursive step, suppose that for some \(n \geq 1\) we have constructed the subgroups \(G_m\) and the enumerations \((\varphi_{m,j})_{j=1}^{\infty}\) of \(\bigcup_{k=1}^{\infty} \operatorname{Hom}(G_m, \mathcal{B}_k)\) for all \(1 \leq m \leq n\). Let \(j_n:= \#\{ i \in \mathbb{N} : 1 \leq i \leq n \text{ and } \ell_i = \ell_n\}\). Consider the map \(\psi_n : G_n' \to S_k\) given by the composition of the following maps: \[\begin{tikzcd} {G_n'} & {G_n} & {G_{\ell_n}} & {\mathcal{B}_k} & {S_k} \arrow[hook, from=1-1, to=1-2] \arrow[hook, from=1-2, to=1-3] \arrow["{\varphi_{\ell_n,j_n}}", from=1-3, to=1-4] \arrow["\tau", from=1-4, to=1-5] \end{tikzcd}\] Here, the first two maps are inclusions, and \(\tau\) is the canonical map (note that \(k\) depends on \(\ell_n\) and \(j_n\)). Since \(G_n\) is a free group on two generators, its commutator subgroup \(G_n'\) is also free of countably infinite rank. Thus if \(H_{n+1} := \ker \psi_n\), then \(H_{n+1}\) must be a countably infinite free group, since \([G_n' : H_{n +1}] = |S_k| = k!\) is finite. Moreover, the rank of \(H_{n+1}\) must be at least two; otherwise, \(H_{n+1}\) would be contained as a finite index subgroup of a rank two subgroup of \(G_{n}'\), which by the Nielsen-Schreier Theorem would imply that \([G_n' : H_{n+1}] = 0\), a contradiction. Thus, we can choose a rank two subgroup of \(H_{n+1}\), and define \(G_{n+1}\) to be this subgroup. Then \(G_{n+1}\) is finitely generated, so the set \(\bigcup_{k=1}^{\infty} \operatorname{Hom}(G_{n+1}, \mathcal{B}_k)\) is countable, and we can choose an enumeration \((\varphi_{n+1,j})_{j=1}^{\infty}\) for this set. This completes the recursion.
Now notice that we have constructed a sequence \(G_1 \supseteq G_2 \supseteq G_3 \supseteq \cdots\) of nested subgroups, each of which is a free group on two generators (so property (1) is satisfied). Property (2) is also satisfied by construction, since \(G_{n+1}\) is contained in the kernel of a map with domain \(G_{n}'\). Finally, given a group homomorphism \(\varphi : G_n \to \mathcal{B}_k\) for some \(n\) and some \(k\), there is a positive integer \(j\) such that \(\varphi = \varphi_{n,j}\). Since every positive integer appears in the sequence \((\ell_i)\) infinitely often, there is some smallest positive integer \(i_0\) such that \(j = \#\{ i \in \mathbb{N} : 1 \leq i \leq i_0 : \ell_i = n\}\); note that \(n = \ell_{i_0} \leq i_0\). Then by construction, at the \(i_0^{\text{th}}\) step of the recursion, we produce a group \(G_{i_0 + 1}\) which is contained in the kernel of the map \(\tau \circ \varphi_{\ell_{i_0}, j_{i_0}} = \tau \circ \varphi_{n, j} = \tau \circ \varphi\), as desired. This shows property (3) is satisfied.
This nested sequence of subgroups defines a pro-group \(\underline{G}\) given by the inverse system of inclusions \[\begin{tikzcd} G_1 & G_2 & G_3 & \cdots \arrow[from=1-2, to=1-1] \arrow[from=1-3, to=1-2] \arrow[from=1-4, to=1-3] \end{tikzcd}\] such that, by property (2), each bonding map factors through the commutator subgroup of its range. In particular, this implies that the abelianization of \(\underline{G}\) is the trivial group. Moreover, property (3) guarantees that the pro-group \(\underline{G}\) satisfies \((**_k)\), and hence \((*_k)\), for every \(k\). Finally, applying Lemma 30, we can find a 1-dimensional continuum \(X\) having \(\underline{\pi}_1(X,x_0) \cong \underline{G}\), which can be realized as an inverse limit of wedges of two circles. Then \(\underline{\pi}_1(X,x_0)\) satisfies \((*_k)\) for every \(k\), and it is non-abelian, since the abelianization of \(\underline{G}\) is trivial (but \(\underline{G}\) is non-trivial, since the bonding maps are all inclusions of non-zero subgroups). Moreover, because \(\mathsf{Ab}(\underline{G}) = 0\), we have that \(\underline{H}_1(X) \cong \mathsf{Ab}(\underline{\pi}_1(X, x_0)) \cong \mathsf{Ab}(\underline{G}) = 0\), so by the UCT, \[\check{H}^1(X) = \varinjlim \underline{H}^1(X) \cong \varinjlim \operatorname{Hom}(\underline{H}_1(X), \mathbb{Z}) = 0\] and hence \(X\) is acyclic.
Our next results concern using more easily computable invariants, such as the first Čech cohomology group \(\check{H}^1(X)\) discussed above, and the homology pro-group \(\underline{H}_1(X)\), in order to discern if low degree polynomials have continuous approximate solutions. For these results, we will need the concept of \(m\)-divisibility for a pro-group.
Definition 39. A pro-group \(\underline{G} = (G_\alpha, p_\alpha^{\alpha'}, A)\) is called \(m\)-divisible if for any \(\alpha \in
A\) there exists a \(\beta \geq \alpha\) so that for any \(h \in G_\beta\) there exists an element \(g \in G_\alpha\) so that \(p_\alpha^\beta(h) = g^m\).
In particular, if \(\underline{G}\) is \(m\)-divisible, so is any quotient of \(\underline{G}\) (such as it’s abelianization), and any homomorphism \(\underline{\varphi}: \underline{G} \to \mathbb{Z}^k\) has its image in the subgroup \(\bigcap_{a \geq 1}m^a\mathbb{Z}^k\).
Lemma 40. Consider a pro-group \(\underline{G}\) and its abelianization \(\operatorname{Pro}(\mathsf{Ab})(\underline{G})\). If \(W\) is a solvable group of exponent \(m\) and \(\operatorname{Pro}(\mathsf{Ab})(\underline{G})\) is \(m\)-divisible, then any morphism \(\underline{\varphi}: \underline{G} \to W\) is trivial.
Proof. Since \(W\) is a finite solvable group, it has a subnormal series \[1 = W_0 \trianglelefteq W_1 \trianglelefteq \cdots \trianglelefteq W_{n - 1} \trianglelefteq W_n = W\] such that \(W_k/W_{k - 1}\) is abelian for each \(1 \leq k\). We shall prove the statement by induction on the length \(n\) of this series. If \(n = 0\) the statement is trivial as \(W\) is already the trivial group.
Now assuming that the statement is true for groups with such a series of length \(n - 1\), take \(W\) to have such a series of length \(n\). Then \(W/W_{n - 1} = W_n/W_{n - 1}\) is an abelian group of exponent (dividing) \(m\). Hence if \[\varphi_\alpha: G_\alpha \to W\] represents \(\underline{\varphi}\), then the composition with the quotient \(W \to W/W_{n - 1}\) factors through \(\mathsf{Ab}(G_\alpha)\), and using the definition of \(\operatorname{Pro}(\mathsf{Ab})(\underline{G})\) being \(m\)-divisible we can find \(\beta \geq \alpha\) so that the image of \(\mathsf{Ab}(G_\beta)\) in \(\mathsf{Ab}(G_\alpha)\) consists only of \(m^{\text{th}}\) powers of elements in \(\mathsf{Ab}(G_\alpha)\). Thus the composition \[\begin{tikzcd} {G_\beta} & {\mathsf{Ab}(G_\beta)} \\ {G_\alpha} & {\mathsf{Ab}(G_\alpha)} & {W/W_{n - 1}} \arrow[from=1-1, to=1-2] \arrow[from=1-1, to=2-1] \arrow[from=1-2, to=2-2] \arrow["0", from=1-2, to=2-3] \arrow[from=2-1, to=2-2] \arrow[from=2-2, to=2-3] \end{tikzcd}\] is trivial. Hence the map from \(G_\beta\) to \(W/W_{n - 1}\) is trivial, and so \(\varphi_\beta: G_\beta \to W\) must actually have its image land in \(W_{n - 1}\). But \(W_{n - 1}\) has a shorter series and therefore the map \(\underline{\varphi}\) is trivial. ◻
Theorem 41. Consider a continuum \(X\) and an integer \(1 \leq n \leq 4\). If \(\underline{H}_1(X)\) is \(n!\)-divisible, then all polynomials \(P: X \to B_n\) are completely solvable.
Proof. Picking a basepoint \(x_0 \in X\), by Theorem 32 we need to show \(\underline{\pi}_1(X, x_0)\) satisfies \((**_n)\), meaning that any \(\underline{\varphi}: \underline{\pi}_1(X, x_0) \to \mathcal{B}_n\) is such that \(\underline{\psi} = \tau \circ \underline{\varphi}\) is trivial, where \(\tau : \mathcal{B}_n \to S_n\) denotes the canonical map. But since \(n \leq 4\) we have that \(\underline{\psi}: \underline{\pi}_1(X, x_0) \to S_n\) is a morphism into a solvable group of order \(n!\), so By Lemma 40 the abelianization \(\underline{H}_1(X)\) of \(\underline{\pi}_1(X, x_0)\) being \(n!\)-divisible guarantees any morphism into \(S_n\) is trivial. ◻
Next we show that for polynomials of degree less than \(3\), we can replace divisibility of \(\underline{H}_1(X)\) with divisibility of \(\check{H}^1(X)\), but not for degree \(n = 4\) polynomials. For these proofs we need to know more about the structure of the braid group \(\mathcal{B}_n\). Recall that \(\mathcal{B}_1\) is defined to be the trivial group, while in general, the braid group on \(n\) strands (for \(n\geq 2\)) has a presentation given by generators \(\sigma_1, \dots, \sigma_{n - 1}\) and relations \[\begin{align} \sigma_i \sigma_j & = \sigma_j \sigma_i \quad \text{for} \;\text{1} \leq i \leq j - 2 \leq n - 3 \\ \sigma_i \sigma_{i + 1} \sigma_i & = \sigma_{i + 1} \sigma_i \sigma_{i + 1} \quad \text{for} \;\text{1} \leq i \leq n - 2 \end{align}\] Thus \(\mathcal{B}_2\) has a single generator and no relations, so \(\mathcal{B}_2 \cong \mathbb{Z}\). It is not too hard to see that these relationships imply that the abelianization \(\mathsf{Ab}(\mathcal{B}_n)\) is \(\mathbb{Z}\) with the abelianization homomorphism given by \(\sigma_i \mapsto 1\) for each \(1 \leq i \leq n - 1\). For our proofs we need to know a bit more about the derived series of \(\mathcal{B}_n\), which is given to us by [5]:
For \(\mathcal{B}_3\): the commutator subgroup \(\mathcal{B}_3'\) is free, generated by the two elements \[u = \sigma_2 \sigma_1^{-1}, \quad v = \sigma_1 \sigma_2 \sigma_1^{-2}\]
For \(\mathcal{B}_4\): the commutator subgroup \(\mathcal{B}_4'\) has its presentation given by four generators \[u, \quad v, \quad a = \sigma_3 \sigma_1^{-1}, \quad b = uau^{-1}\] and relations \[uau^{-1} = b, \quad ubu^{-1} = b^2a^{-1}b, \quad vav^{-1} = a^{-1}b, \quad vbv^{-1} = (a^{-1}b)^3 a^{-2}b\] where \(u\) and \(v\) are as above. The additional relations \[u^{-1}au = ab^{-1}a^2, \quad u^{-1}bu = a, \quad v^{-1}av = ab^{-1}a^3, \quad v^{-1}bv = ab^{-1}a^4\] hold, making the subgroup \(T \subseteq\mathcal{B}_4'\) generated by \(a\) and \(b\) a normal subgroup. In fact \(T\) is freely generated by \(a\) and \(b\), and the quotient \(\mathcal{B}_4'/T\) is a free group generated by the images of \(u\) and \(v\).
For \(\mathcal{B}_n\) with \(n \geq 5\): the commutator subgroup \(\mathcal{B}_5'\) is perfect, meaning the second commutator subgroup \(\mathcal{B}_5''\) is equal to the first \(\mathcal{B}_5'\).
Recall from Section 2.1 that we have an action \(\star: \mathbb{C}^\times \times B_n \to B_n\) given by scaling roots of a polynomial at a point by a non-zero complex number. Then if we consider a continuum \(X\), a degree \(n\) polynomial \(P: X \to B_n\), and a non-vanishing function \(f: X \to \mathbb{C}^\times\), we can define \(f \star P\) by applying this action pointwise. Observe that finding solutions to \(P\) and \(f\star P\) are equivalent problems. In particular, if we can find a non-vanishing function \(f\) such that the discriminant of \(f\star P\) is constant and equal to \(1\), then we can find solutions to \(P\) by finding solutions to \(f\star P\). This is useful, because if \(f\star P :X \to B_n\) has constant discriminant equal to \(1\), then its image lies in the subspace \(B_n'\) of \(B_n\), which has fundamental group \(\mathcal{B}_n'\) (see Section 2), so the induced map \(\underline{\pi}_1 (f\star P) : \underline{\pi}_1(X,x_0) \to \mathcal{B}_n\) on fundamental pro-groups has its image lying in \(\mathcal{B}_n'\). This is helpful in the context of Corollary 31 for finding solutions to \(f\star P\) (and hence solutions of \(P\)).
To find such an \(f\), recall that \(\Delta(f \star P) = f^{n(n - 1)} \Delta(P)\), so we need to make \(f\) an \((n(n-1))^{\text{th}}\) root of the function \(\frac{1}{\Delta(P)} : X \to \mathbb{C}^\times\). Such an \(f\) exists if \(\check{H}^1(X)\) is \(n(n-1)\)-divisible, by Corollary 36.
With the above discussion, we can now prove that for quadratic and cubic polynomials, \(2\)- and \(3\)-divisibility of \(\check{H}^1(X)\) is enough to find solutions.
Theorem 42. For \(n = 2\) or \(n = 3\), a continuum \(X\) has an \(n!\)-divisible \(\check{H}^1(X)\) if and only if any polynomial \(P: X \to B_n\) is completely solvable.
Proof. For the backwards direction, we can get \(2\)- or \(3\)-divisibility of \(\check{H}^1(X)\) by using the fact that we can find roots of non-vanishing functions (Corollary 36), so let’s look at the forwards direction.
The case \(n = 2\) is simple. As we argued above, \(\check{H}^1(X)\) being \(2\)-divisible ensures that we only need to worry about homomorphisms \(\underline{\varphi}: \underline{\pi}_1(X, x_0) \to \mathcal{B}_2\) whose image lands in \(\mathcal{B}_2'\) which is the trivial group. But these definitely have their image in \(N_2\), so \(\underline{\pi}_1(X, x_0)\) satisfies \((**_2)\).
The case \(n = 3\) is trickier. Consider some \(\underline{\varphi}: \underline{\pi}_1(X, x_0) \to \mathcal{B}_3\), the image of which we may similarly assume lands in \(\mathcal{B}_3'\), as \(\check{H}^1(X)\) is \(6\)-divisible. Since the image of the commutator subgroup \(\mathcal{B}_3'\) under the canonical map \(\tau: \mathcal{B}_3 \to S_3\) lies in the abelian subgroup \(S_3' = A_3 \cong \mathbb{Z}/3\mathbb{Z}\), it follows that \(\tau|_{\mathcal{B}_3'}\) factors through the abelianization of \(\mathcal{B}_3'\). But by the discussion above, \(\mathcal{B}_3'\) is freely generated by two elements, so the abelianization of \(\mathcal{B}_3'\) is \(\mathbb{Z}^2\). In total we get the diagram: \[\begin{tikzcd} & {3\mathbb{Z}^2} \\ {\underline{\pi}_1(X, x_0)} & {\mathbb{Z}^2} & {A_3 \cong \mathbb{Z}/3\mathbb{Z}} \\ & {\mathcal{B}_3'} \arrow[hook, from=1-2, to=2-2] \arrow["0", from=1-2, to=2-3] \arrow[dashed, from=2-1, to=1-2] \arrow["{\underline{\varphi}}", from=2-1, to=3-2] \arrow[from=2-2, to=2-3] \arrow[from=3-2, to=2-2] \arrow["{\tau|_{\mathcal{B}_3'}}"', from=3-2, to=2-3] \end{tikzcd}\] By our assumption that the dual group \[\check{H}^1(X) = \operatorname{Hom}(\underline{\pi}_1(X, x_0), \mathbb{Z})\] is \(3\)-divisible, by Proposition 35 we see that the the map \(\underline{\pi}_1(X, x_0) \to \mathbb{Z}^2\) given by \(\underline{\varphi}\) and then abelianization, has its image in \(3\mathbb{Z}^2\). Therefore \(\tau \circ \underline{\varphi}\) is the trivial map, which in total shows that \(\underline{\pi}_1(X, x_0)\) satisfies \((**_3)\) as required. ◻
Here are some counterexamples in the cases \(n = 4\) and \(n \geq 5\). First the simpler case, \(n = 5\) when \(A_n\) is not solvable (and moreover perfect).
Lemma 43. Given any integer \(n \geq 5\) there exists a pro-group \(\underline{G^n}\) given by an inverse system of free groups on two generators \[\begin{tikzcd} {\mathbb{Z}* \mathbb{Z}} & {\mathbb{Z}* \mathbb{Z}} & {\mathbb{Z}* \mathbb{Z}} & \cdots \arrow["f_n"', from=1-2, to=1-1] \arrow["f_n"', from=1-3, to=1-2] \arrow[from=1-4, to=1-3] \end{tikzcd}\] that does not satisfy \((*_n)\), but has \(\mathsf{Ab}(\underline{G^n}) = 0\).
Proof. For \(n \geq 5\) we know that the group \(A_n\) is perfect and is generated by two elements \(a, b \in A_n\). In particular we have a surjective morphism \(\psi: \mathbb{Z}* \mathbb{Z}\to A_n\) by sending the free generators \(x, y\) of \(\mathbb{Z}* \mathbb{Z}\) to \(a\) and \(b\) respectively. But now we get a commutative diagram \[\begin{tikzcd} {\mathbb{Z}* \mathbb{Z}} & {(\mathbb{Z}* \mathbb{Z})'} \\ {A_n} & {A_n'} \arrow["\psi"', from=1-1, to=2-1] \arrow[from=1-2, to=1-1] \arrow[from=1-2, to=2-2] \arrow["\operatorname{id}"', from=2-2, to=2-1] \end{tikzcd}\] obtained by restricting to the commutator subgroups, where the bottom map is the identity (as \(A_n'\) is all of \(A_n\)). Therefore the restriction of \(\psi\) to \((\mathbb{Z}* \mathbb{Z})'\) is still surjective, so we can find \(x', y' \in (\mathbb{Z}* \mathbb{Z})'\) such that \(\psi(x') = a\) and \(\psi(y') = b\). We take \(f_n: \mathbb{Z}* \mathbb{Z}\to \mathbb{Z}* \mathbb{Z}\) to be the endomorphism sending \(x\) to \(x'\) and \(y\) to \(y'\). Then by construction \(\psi \circ f_n = \psi\), and hence \(\psi \circ f_n^{\circ m} = \psi\) for any natural number \(m\).
To define \(\varphi\) we lift \(\psi: \mathbb{Z}* \mathbb{Z}\to A_n\) along the natural map \(\tau|_{\mathcal{B}_n'}: \mathcal{B}_n' \to A_n\), which is possible as \(\tau\) (and hence \(\tau|_{\mathcal{B}_n'}\)) is surjective and \(\mathbb{Z}* \mathbb{Z}\) is free. Now we take \(\underline{G^n}\) to be the inverse system where all of the structure maps are \(f_n\), and take the morphism \(\underline{\varphi}: \underline{G^n} \to \mathcal{B}_n' \subseteq\mathcal{B}_n\) given by \(\varphi\) from the first object of the inverse system.
If the image of \(\underline{\varphi}\) were contained in some \(M_{n,i}\), then we should be able to factor \(\varphi\) through \(M_{n,i}\) by going far enough up the inverse system, meaning \(\varphi\circ f_n^{\circ m}\) would have to have its image in \(M_{n,i}\) for some large enough \(m\). However \(M_{n,i}\) is exactly the preimage under \(\tau\) of the subgroup of \(S_n\) consisting of permutations that fix the \(i\)th element, so it is enough to check that \(\tau \circ \varphi\circ f_n^{\circ m}\) never lies in these subgroups. But \[\tau \circ \varphi\circ f_n^{\circ m} = \psi \circ f_n^{\circ m} = \psi\] which has an image of \(A_n\), and no element is fixed by all permutations in this subgroup. Therefore \(\underline{G^n}\) does not satisfy \((*_n)\).
Lastly we need to argue that \(\mathsf{Ab}(\underline{G^n}) = 0\), but this is immediate as \(f_n: \mathbb{Z}* \mathbb{Z}\to \mathbb{Z}* \mathbb{Z}\) has its image landing in the commutator subgroup, so \(f_n\) becomes the zero map after abelianizing and hence the inverse system consists of a sequence of zero maps. ◻
Example 3. As in Lemma 30 we can find a \(1\)-dimensional continuum \(X\) which has \(\underline{\pi}_1(X, x_0) \cong \underline{G^n}\) from Lemma 43 by taking the inverse limit of wedges of two circles \[\begin{tikzcd} {S^1 \vee S^1} & {S^1 \vee S^1} & {S^1 \vee S^1} & \cdots \arrow["{F_n}"', from=1-2, to=1-1] \arrow["{F_n}"', from=1-3, to=1-2] \arrow[from=1-4, to=1-3] \end{tikzcd}\] using a pointed map \(F_n\) that realizes the morphism \(f_n\). This continuum \(X\) carries a polynomial \(P: X \to B_n\) with no solutions by Theorem 32, because \(\underline{\pi}_1(X)\) does not satisfy \((*_n)\). On the other hand, this \(X\) has \(\underline{H}_1(X) \cong \mathsf{Ab}(\underline{\pi}_1(X, x_0)) \cong \mathsf{Ab}(\underline{G^n}) = 0\) so by the UCT, \[\check{H}^1(X) = \varinjlim \underline{H}^1(X) \cong \varinjlim \operatorname{Hom}(\underline{H}_1(X), \mathbb{Z}) = 0\] showing in particular that \(\check{H}^1(X)\) is divisible. Note that by [2], this continuum \(X\) is an example of a continuum that admits approximate \(m^{\text{th}}\) roots for every \(m \geq 1\), but does not admit approximate continuous solutions to some degree \(n\) polynomial.
Remark 44. Note that an acyclic one-dimensional continuum will always satisfy \((*_4)\). Indeed, if \(X\) is one-dimensional, then its shape is given by an inverse sequence of wedges of circles, and hence the homology pro-group \(\underline{H}_1(X)\) is an inverse limit of direct sums of copies of \(\mathbb{Z}\). If additionally \(X\) is acyclic, then by the UCT, \(\varinjlim \operatorname{Hom}(\underline{H}_1(X), \mathbb{Z}) = \check{H}^1(X) = 0\), which implies that \(\underline{H}_1(X)\) is \(m\)-divisible for all \(m \geq 1\). In particular, it is \(4!\)-divisible, and hence Theorem 41 applies.
We now move to the more difficult case of find an acyclic two-dimensional continuum whose fundamental pro-group does not satisfy \((*_4)\). We start with a technical lemma that will help us build the appropriate pro-group.
Lemma 45. Consider two elements \(U, V \in \operatorname{SL}_2(\mathbb{Z})\). The subgroup they generate in \(\operatorname{SL}_2(\mathbb{Z})\) is freely generated by them if and only if their images in \(\operatorname{PSL}_2(\mathbb{Z})\) also freely generate a subgroup. In this case, the sum of the ranges \[\operatorname{Im}(U - \operatorname{id}) + \operatorname{Im}(V - \operatorname{id}) \subseteq\mathbb{Z}^2\] is a rank \(2\) subgroup.
Proof. Let us denote the quotient map by \(f:\operatorname{SL}_2(\mathbb{Z}) \to \operatorname{PSL}_2(\mathbb{Z})\), which has a kernel of \(\{\pm \operatorname{id}\}\). Let’s start with the equivalence between generating a free subgroup of \(\operatorname{SL}_2(\mathbb{Z})\) and \(\operatorname{PSL}_2(\mathbb{Z})\).
If \(U, V\) are free generators of their subgroup \(\left\langle U, V \right\rangle\), then we cannot have \(-\operatorname{id}\) as an element of \(\left\langle U, V \right\rangle\) as this subgroup is torsion-free. Therefore \(\left\langle U, V \right\rangle \cap \ker(f) = \{\operatorname{id}\}\) meaning that \(f\) restricts to be injective on \(\left\langle U, V \right\rangle\) as required. Conversely, if \(f(U), f(V)\) freely generate a subgroup in \(\operatorname{PSL}_2(\mathbb{Z})\) it cannot be that there are any relations between \(U\) and \(V\), as these would give relations between \(f(U)\) and \(f(V)\).
Next we show that under these conditions, the sum of the ranges \(\operatorname{Im}(U - \operatorname{id}) + \operatorname{Im}(V - \operatorname{id}) \subseteq\mathbb{Z}^2\) is a rank \(2\) subgroup. First note that it cannot be that \(U = \operatorname{id}\) or \(V = \operatorname{id}\) if they freely generate a subgroup, therefore \(\operatorname{Im}(U - \operatorname{id})\) and \(\operatorname{Im}(V - \operatorname{id})\) are both at least rank \(1\). Then if \(\operatorname{Im}(U - \operatorname{id}) + \operatorname{Im}(V - \operatorname{id})\) is not rank \(2\), it must be that \(\operatorname{Im}(U - \operatorname{id})\) and \(\operatorname{Im}(V - \operatorname{id})\) are subsets of a common \(\mathbb{Z}w_1\), where by taking out common factors we may assume \(\mathbb{Z}w_1 = (\mathbb{Q}w_1) \cap \mathbb{Z}^2\). In particular \(w_1\) is an eigenvector for both \(U\) and \(V\).
Completing \(w_1\) to a basis \(\{w_1, w_2\}\) for \(\mathbb{Z}^2\), we see that in this basis \(f(U)\) and \(f(V)\) are matrices of the form \[\begin{bmatrix} 1 & * \\ 0 & 1\end{bmatrix}\] inside of \(\operatorname{PSL}_2(\mathbb{Z})\) which commute, and hence cannot freely generate a subgroup. ◻
Lemma 46. There exists a nested sequence of subgroups \[\mathcal{B}_4' = G_0 \supseteq G_1 \supseteq G_2 \supseteq\cdots\] such that the pro-group \(\underline{G}\) given by the inverse system of inclusions \[\begin{tikzcd} G_0 & G_1 & G_2 & \cdots \arrow[from=1-2, to=1-1] \arrow[from=1-3, to=1-2] \arrow[from=1-4, to=1-3] \end{tikzcd}\] does not satisfy \((*_4)\), but has trivial dual group \(\varinjlim \operatorname{Hom}(\underline{G}, \mathbb{Z}) = 0\).
Proof. Recall that \(\mathcal{B}_4'\) is generated by four elements: \(u, v, a, b\), such that \(a\) and \(b\) freely generate a normal subgroup \(T\) of \(\mathcal{B}_4'\). The elements \(u\) and \(v\) also freely generate a subgroup \(J\). We define the pro-group \(\underline{G}\) by picking an injection \(\mathbb{Z}* \mathbb{Z}\to (\mathbb{Z}* \mathbb{Z})'\) such as the subgroup generated by \(xyx^{-1}y^{-1}\) and \(xy^2x^{-1}y^{-2}\), with which we can recursively define \[\begin{align} u_0 = u \quad & v_0 = v \\ u_n = u_{n - 1}v_{n - 1}u_{n - 1}^{-1}v_{n - 1}^{-1} \quad & v_n = u_{n - 1}v_{n - 1}^2u_{n - 1}^{-1}v_{n - 1}^{-2} \quad \text{for} \;n \geq 1 \end{align}\] to get the subgroups \(J_n = \left\langle u_n, v_n \right\rangle \subseteq J\) and \(G_n = TJ_n \subseteq\mathcal{B}_4'\).
First note that the standard inclusions \(G_n \hookrightarrow \mathcal{B}_4\) induce a morphism \(\underline{\varphi}: \underline{G} \to \mathcal{B}_4\) which does not have its image in one of the \(M_{4, i}\). This is because all of the subgroups \(G_n\) contain \(T = \left\langle a, b \right\rangle = \left\langle \sigma_3 \sigma_1^{-1}, \sigma_2 \sigma_1^{-1}\sigma_3 \sigma_2^{-1} \right\rangle\) whose image under the canonical map \(\mathcal{B}_4 \to S_4\) is the subgroup \(\left\langle (12)(34), (13)(24) \right\rangle\) of \(S_4\). Since no \(i \in \{1,2,3,4\}\) is a common fixed point of these permutations, \(T\) (and hence none of the \(G_n\)) lie in any \(M_{4, i}\), so \(\underline{G}\) doesn’t satisfy \((*_4)\).
We want to understand the groups \(\operatorname{Hom}(G_n, \mathbb{Z})\), which comes down to understanding how \(u_n\) and \(v_n\) act on \(a\) and \(b\). Since \(T\) is a normal subgroup, conjugation by an element of \(J\) descends to an action on the abelianization \(\mathsf{Ab}(T) \cong \mathbb{Z}^2\). By using the images of \(a\) and \(b\) as a basis for \(\mathsf{Ab}(T)\), we get a map \(\alpha: J \to \operatorname{Aut}(\mathsf{Ab}(T)) \cong \operatorname{GL}_2(\mathbb{Z})\) where \(\alpha(x) \in \operatorname{Aut}(\mathsf{Ab}(T))\) is the automorphism defined by \[\alpha(x)(y T') = xyx^{-1} T'.\] Using the relations for \(\mathcal{B}_4'\) we stated above we find \[uau^{-1} = b \mod T'\;\;\;\;\;\; \text{ and } \;\;\; \quad ubu^{-1} = a^{-1}b^3 \mod T'\] so the matrix \(U := \alpha(u)\) in the \(a,b\) basis is given by \[\begin{bmatrix}0 & -1 \\ 1 & 3\end{bmatrix}.\] Similarly for conjugation by \(v\), \[vav^{-1} = a^{-1}b \mod T' \;\;\;\;\;\; \text{ and } \;\;\; \quad vbv^{-1} = a^{-5}b^4 \mod T'\] so the matrix \(V := \alpha(v)\) is given by \[\begin{bmatrix}-1 & -5 \\ 1 & 4\end{bmatrix}.\] Note this computation shows the image of \(\alpha\) lies in \(\operatorname{SL}_2(\mathbb{Z})\). We claim that \(U\) and \(V\) freely generate a subgroup of \(\operatorname{SL}_2(\mathbb{Z})\), and we will use this to prove that \(\varinjlim \operatorname{Hom}(\underline{G}, \mathbb{Z}) = 0\). Note that if \(U_0 = U\) and \(V_0 = V\) freely generate a subgroup, then by induction so do \[\begin{align} U_n & = \alpha(u_n) = U_{n - 1}V_{n - 1}U_{n - 1}^{-1}V_{n - 1}^{-1} \quad \text{ and }\\ V_n & = \alpha(v_n) = U_{n - 1}V_{n - 1}^2U_{n - 1}^{-1}V_{n - 1}^{-2}. \end{align}\] Before proving the claim, let us show that all of the structure maps in \(\operatorname{Hom}(\underline{G}, \mathbb{Z})\) are zero (and hence the limit is zero). By definition the subgroup \(G_n\) is generated by \(T\) and the elements \(u_n\) and \(v_n\). Therefore to understand a structure map \[\begin{tikzcd} {\operatorname{Hom}(G_{n - 1}, \mathbb{Z})} & {\operatorname{Hom}(G_n, \mathbb{Z})} \arrow[from=1-1, to=1-2] \end{tikzcd}\] induced by the inclusion of subgroups, we need to take a morphism \(f: G_{n - 1} \to \mathbb{Z}\) and compute its value on \(T\) and \(u_n\) and \(v_n\). We see that \(f\) automatically vanishes on \(u_n\) and \(v_n\) as they are commutators of elements in \(G_{n - 1}\). Since \(f\) maps into the abelian \(\mathbb{Z}\), its restriction to \(T\) factors through \(\mathsf{Ab}(T) \cong \mathbb{Z}^2\), and to show it is the zero map it is enough to show that \(\tilde{f}: \mathsf{Ab}(T) \cong \mathbb{Z}^2 \to \mathbb{Z}\) has a rank \(2\) subgroup in its kernel. Given any \(x \in J_n\) and \(y \in T\) we have \[\tilde{f}(y T') = \tilde{f}(x T') + \tilde{f}(y T') - \tilde{f}(x T') = \tilde{f}(xyx^{-1} T') = \tilde{f}(\alpha(x)(y T'))\] showing that \(\tilde{f}\) is zero on any elements \((\alpha(x) - \operatorname{id})y\). In particular the images of both \(U_n - \operatorname{id}\) and \(V_n - \operatorname{id}\) are in the kernel of \(\tilde{f}\). Then by the claim we know \(U_n\) and \(V_n\) freely generate a subgroup of \(\operatorname{SL}_2(\mathbb{Z})\), so by Lemma 45 the sum of these images is a rank \(2\) subgroup of \(\mathbb{Z}^2\).
Finally we need to show that \[U = \begin{bmatrix}0 & -1 \\ 1 & 3\end{bmatrix} \quad \text{ and } \quad V = \begin{bmatrix}-1 & -5 \\ 1 & 4\end{bmatrix}\] freely generate a subgroup of \(\operatorname{SL}_2(\mathbb{Z})\), which by Lemma 45 is the same as showing that their images in \(\operatorname{PSL}_2(\mathbb{Z})\) freely generate a subgroup. It is a standard result that \(\operatorname{PSL}_2(\mathbb{Z})\) has the following presentation (see [20]): \[\operatorname{PSL}_2(\mathbb{Z}) = \left\langle S, Q \;| \;S^2 = 1 , \quad Q^3 = 1 \right\rangle \cong \mathbb{Z}/2\mathbb{Z}* \mathbb{Z}/3\mathbb{Z}\] where the matrices \(S\) and \(Q\) are given by \[S = \begin{bmatrix}0 & -1 \\ 1 & 0\end{bmatrix} \quad Q = \begin{bmatrix}1 & -1 \\ 1 & 0\end{bmatrix}\] and our matrices are \(U = S(QS)^3\) and \(V = SQ^{-1}SQU\). We shall compute the subgroup \(\left\langle U, V \right\rangle\) by using covering space theory, which will along the way show that \(\left\langle U, V \right\rangle\) is the commutator subgroup of \(\operatorname{PSL}_2(\mathbb{Z})\).
To start note that when abelianizing \(\operatorname{PSL}_2(\mathbb{Z})\) the element \(U\) is congruent to \(S^4Q^3 = 1\) and \(V\) is congruent to \(S^2U = 1\), so \(\left\langle U, V \right\rangle\) is a subgroup of the commutator subgroup of \(\operatorname{PSL}_2(\mathbb{Z})\). We can form a CW complex \(Y\) with a basepoint \(y_0\), two \(1\)-cells which we label \(S\) and \(Q\), and two \(2\)-cells implementing the relations \(S^2 = 1\) and \(Q^3 = 1\). Thus \(\pi_1(Y, y_0) \cong \operatorname{PSL}_2(\mathbb{Z})\) via the same presentation as above, and we can consider the covering space \((\widetilde{Y}, \widetilde{y_0})\) corresponding to the commutator subgroup of \(\pi_1(Y, y_0)\).
The covering map \((\widetilde{Y}, \widetilde{y_0})\) to \((Y, y_0)\) is a \(6\)-fold covering as the abelianization of \(\operatorname{PSL}_2(\mathbb{Z})\) is \(\mathbb{Z}/2\mathbb{Z}\oplus \mathbb{Z}/3\mathbb{Z}\), an order \(6\) group. Giving \((\widetilde{Y}, \widetilde{y_0})\) the CW structure induced by the covering map, the \(1\)-skeleton of \((\widetilde{Y}, \widetilde{y_0})\) is as in Figure 1, as the Deck transformations for this space are \(\mathbb{Z}/2\mathbb{Z}\oplus \mathbb{Z}/3\mathbb{Z}\). Since \(\left\langle U, V \right\rangle\) is a subgroup of the commutator subgroup, the two elements \(U\) and \(V\) still represent loops in \(\widetilde{Y}\) based at \(\widetilde{y_0}\). Collapsing the six \(2\)-cells associated to \(Q^3 = 1\) to two points (one point by identifying \(\widetilde{y_0}\), \(Q\widetilde{y_0}\), \(Q^2\widetilde{y_0}\) and the other by identifying \(S\widetilde{y_0}\), \(QS\widetilde{y_0}\), \(Q^2S\widetilde{y_0}\)) and collapsing the six \(2\)-cells associated to \(S^2 = 1\) to three line segments (one between \(\widetilde{y_0}\) and \(S\widetilde{y_0}\), another between \(Q\widetilde{y_0}\) and \(SQ\widetilde{y_0}\), and the final between \(Q^2\widetilde{y_0}\) and \(SQ^2\widetilde{y_0}\)) we see that \((\widetilde{Y}, \widetilde{y_0})\) is homotopy equivalent to the theta space and hence has \(\pi_1(\widetilde{Y}, \widetilde{y_0}) \cong \mathbb{Z}* \mathbb{Z}\). Moreover, following the loops \(U\) and \(V\) along this homotopy equivalence, we see that they end up as generators for the fundamental group of the theta space (see Figure 2), and hence \(\left\langle U, V \right\rangle = \pi_1(\widetilde{Y}, \widetilde{y_0}) \cong \mathbb{Z}* \mathbb{Z}\) with \(U\) and \(V\) as generators as required. ◻
Figure 2: On the left we illustrate the loop on \(\widetilde{Y}\) corresponding to \(U\) that starts at \(\widetilde{y_0}\). On the right we depict this loop once we collapse \(\widetilde{Y}\) to the theta space. If we use \(a\) and \(b\) to denote the loops on the theta space by traversing the respective semi-circles counterclockwise, we see that the loop corresponding to \(U\) is \(ab^{-1}\). A similar calculation shows that \(V\) corresponds to \(bab^{-1}\), and these two loops are freely generate the fundamental group..
Example 4. Applying Lemma 30, we can find a 2-dimensional continuum \(X\) for which the pro-group \(\underline{G}\) in Lemma 46 is realized as \(\underline{\pi}_1(X,x_0)\). Since \(\underline{\pi}_1(X,x_0) \cong \underline{G}\) does not satisfy \((*_4)\), it follows that there is a degree 4 polynomial over \(C(X)\) with no repeated roots that does not possess an exact root. Moreover, since \(\varinjlim \operatorname{Hom}(\underline{G}, \mathbb{Z}) = 0\), it follows by the UCT that \[\check{H}^1(X) = \varinjlim \underline{H}^1(X) \cong \varinjlim \operatorname{Hom}(\underline{H}_1(X), \mathbb{Z}) = 0\] so \(X\) is acyclic.