Hostname: page-component-68c7f8b79f-fc4h8 Total loading time: 0 Render date: 2026-01-15T23:08:30.568Z Has data issue: false hasContentIssue false

Partitioning theorems for sets of semi-Pfaffian sets, with applications

Published online by Cambridge University Press:  13 January 2026

Martin Lotz
Affiliation:
University of Warwick , United Kingdom; E-mail: martin.lotz@warwick.ac.uk
Abhiram Natarajan*
Affiliation:
University of Warwick , United Kingdom
Nicolai Vorobjov
Affiliation:
University of Bath , United Kingdom, and St. Petersburg Department of V. A. Steklov Institute of Mathematics of the Russian Academy of Sciences (PDMI RAS), Russia; E-mail: masnnv@bath.ac.uk, vorobjov@pdmi.ras.ru
*
E-mail: abhiram.natarajan@warwick.ac.uk (Corresponding author)

Abstract

We generalize the seminal polynomial partitioning theorems of Guth and Katz [33, 28] to a set of semi-Pfaffian sets. Specifically, given a set $\Gamma \subseteq \mathbb {R}^n$ of k-dimensional semi-Pfaffian sets, where each $\gamma \in \Gamma $ is defined by a fixed number of Pfaffian functions, and each Pfaffian function is in turn defined with respect to a Pfaffian chain $\vec {q}$ of length r, for any $D \ge 1$, we prove the existence of a polynomial $P \in \mathbb {R}[X_1, \ldots , X_n]$ of degree at most D such that each connected component of $\mathbb {R}^n \setminus Z(P)$ intersects at most $\sim \frac {|\Gamma |}{D^{n - k - r}}$ elements of $\Gamma $. Also, under some mild conditions on $\vec {q}$, for any $D \ge 1$, we prove the existence of a Pfaffian function $P'$ of degree at most D defined with respect to $\vec {q}$, such that each connected component of $\mathbb {R}^n \setminus Z(P')$ intersects at most $\sim \frac {|\Gamma |}{D^{n-k}}$ elements of $\Gamma $. To do so, given a k-dimensional semi-Pfaffian set $\mathcal {X} \subseteq \mathbb {R}^n$, and a polynomial $P \in \mathbb {R}[X_1, \ldots , X_n]$ of degree at most D, we establish a uniform bound on the number of connected components of $\mathbb {R}^n \setminus Z(P)$ that $\mathcal {X}$ intersects; that is, we prove that the number of connected components of $(\mathbb {R}^n \setminus Z(P)) \cap \mathcal {X}$ is at most $\sim D^{k+r}$. Finally, as applications, we derive Pfaffian versions of Szemerédi-Trotter-type theorems, and also prove bounds on the number of joints between Pfaffian curves.

Information

Type
Foundations
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2026. Published by Cambridge University Press

1 Introduction

Incidence combinatorics has recently experienced a surge of activity, largely fueled by the introduction of powerful algebraic techniques (see, for example, [Reference Guth29, Reference Sheffer58]). While traditionally restricted to simple geometric objects such as points and lines, the focus in incidence geometry has shifted towards questions involving higher-dimensional algebraic or semi-algebraic sets. Semi-algebraic sets possess convenient tame topological properties, such as triangulability and stratifiability, which play an important role in incidence questions. As suggested in [Reference Grothendieck27], if one were to look for other collections of subsets of $\mathbb {R}^n$ that also possess the tame topological properties of semi-algebraic sets, one is naturally led to o-minimal geometry, an axiomatic generalization of semi-algebraic geometry.

Having its genesis in model theory, the notion of o-minimality isolates key axioms such that any collection of subsets of $\mathbb {R}^n$ satisfying these axioms – such a collection is called an o-minimal structure, and elements of an o-minimal structure are called definable sets – share the tame topological properties that semi-algebraic sets possess. We refer the reader to [Reference Marker44] and [Reference den Dries65] for background in model theory and o-minimal geometry. There is a growing body of work investigating incidence problems where the involved sets are definable over arbitrary o-minimal structures or distal structures over $\mathbb {R}$ [Reference Basu and Raz9, Reference Chernikov, Galvin and Starchenko15, Reference Chernikov and Starchenko17, Reference Chernikov, Starchenko and Thomas18, Reference Basit, Chernikov, Starchenko, Tao and Tran7, Reference Anderson3, Reference Balsera5, Reference Chernikov, Peterzil and Starchenko16]; see [Reference Scanlon55, §6] for an overview. However, compared to the algebraic and semi-algebraic settings, progress in o-minimal incidence combinatorics has been slow.

A particularly useful algebraic tool in incidence geometry is the celebrated polynomial partitioning theorem, first introduced in groundbreaking work by Guth and Katz [Reference Guth and Katz33] to resolve the distinct distances problem and the joints problem. This method was later generalized in [Reference Guth28] and [Reference Blagojević, Blagojević and Ziegler13]. In the statement of this theorem, and in what follows, we denote by $Z(P_1, \dots , P_s)$ the set of common zeros in $\mathbb {R}^n$ of functions $P_1,\dots ,P_s$ .

Theorem 1.1 (Polynomial partitioning [Reference Guth and Katz33, Reference Guth28])

Let $\Gamma $ be a finite set of k-dimensional real algebraic sets in $\mathbb {R}^n$ , where each $\mathcal {X} \in \Gamma $ is defined by at most m polynomials, each of degree at most $\beta $ . For any integer $D \ge 1$ , there exists a nonzero polynomial $P \in \mathbb {R}[X_1, \ldots , X_n]$ of degree at most D and a constant $C = C(n, m, \beta )$ , such that each connected component of $\mathbb {R}^n \setminus Z(P)$ intersects at most $CD^{k - n}|\Gamma |$ elements of $\Gamma $ .

Theorem 1.1 can be interpreted as follows. It is well known (see Theorem 2.13) that given a polynomial $P \in \mathbb {R}[X_1, \ldots , X_n]$ of degree at most D, $\mathbb {R}^n \setminus Z(P)$ can have at most $\mathcal {O}_n(D^n)$ connected components. It is also known (see Theorem 2.14) that out of these $\mathcal {O}_n(D^n)$ connected components, any k-dimensional real algebraic set $\mathcal {X} \subseteq \mathbb {R}^n$ can intersect at most $\mathcal {O}_{n,\mathcal {X}}(D^k)$ of them.Footnote 1 The polynomial partitioning theorem states that given any finite set $\Gamma $ of k-dimensional real algebraic sets and given any $D \ge 1$ , there exists a polynomial P of degree at most D such that the intersections of algebraic sets with connected components are equidistributed. More specifically, since each $\mathcal {X} \in \Gamma $ can intersect at most $\mathcal {O}_{n, \mathcal {X}}(D^k)$ connected components of $\mathbb {R}^n \setminus Z(P)$ , the total number of intersections between elements of $\Gamma $ and connected components of $\mathbb {R}^n \setminus Z(P)$ is at most $\mathcal {O}_{n,\mathcal {X}}(|\Gamma |D^k)$ ; Theorem 1.1 promises that there exists a polynomial P that ensures these intersections are equidistributed among the $\mathcal {O}_n(D^n)$ connected components of $\mathbb {R}^n \setminus Z(P)$ , i.e., there are at most $\mathcal {O}_{n,\mathcal {X}}(|\Gamma |D^{k-n})$ elements of $\Gamma $ intersecting each connected component of $\mathbb {R}^n \setminus Z(P)$ . While not stated this way, Theorem 1.1 holds even if $\Gamma $ is a finite set of semi-algebraic sets, since any semi-algebraic set is contained in a real algebraic set of the same dimension.

The polynomial partitioning theorem gives us a divide and conquer technique; it allows us to break a problem instance into smaller subinstances, solve each subinstance, and then combine the local solutions to obtain the global answer. The polynomial partitioning technique has been a panacea for a large number of problems in discrete geometry [Reference Guth and Katz33, Reference Kaplan, Matoušek and Sharir36, Reference Sheffer and Zahl59, Reference Tidor, Yu and Zhao64], computational geometry/data structures [Reference Afshani and Cheng1, Reference Agarwal, Aronov, Ezra and Zahl2, Reference Aronov, Ezra and Zahl4], harmonic analysis [Reference Guth30, Reference Guth31], and others.

A long-standing open problem that has stymied progress on incidence questions involving sets definable over arbitrary o-minimal expansions of $\mathbb {R}$ is the unavailability of a polynomial partitioning type result for definable sets; see comments in [Reference Basu and Raz9, Reference Chernikov, Galvin and Starchenko15, Reference Anderson3, Reference Balsera5, Reference Chernikov, Peterzil and Starchenko16]. Needless to say, a polynomial partitioning theorem that works in the o-minimal setting would enable tremendous progress in o-minimal incidence combinatorics and would also provide greatly simplified proofs of existing results.

Remark 1.2 (Difficulty in obtaining an o-minimal polynomial partitioning theorem)

To prove a polynomial partitioning theorem in the o-minimal setting, a key step is to obtain a result of the type of Theorem 2.14 in the case where $\mathcal {X} \subseteq \mathbb {R}^n$ is a k-dimensional definable set in any o-minimal structure. Specifically, given a polynomial $P \in \mathbb {R}[X_1, \ldots , X_n]$ of degree at most D, we would like an upper bound on the number of connected components of $\mathbb {R}^n \setminus Z(P)$ that $\mathcal {X}$ intersects. However, a result of [Reference Basu, Lerario and Natarajan8] suggests that a uniform bound might not always be possible for definable sets in any arbitrary o-minimal structure. Specifically, [Reference Basu, Lerario and Natarajan8] shows that for every $0 \le k \le n-2$ , and every sequence of natural numbers $(a_d)_{d \in \mathbb {N}}$ , there is a regular compact and definable hypersurface $\mathcal {X} \subseteq \mathbb {R} \mathbb {P}^n$ , a subsequence $(a_{d_m})_{m \in \mathbb {N}}$ , and a sequence of homogeneous polynomials $(P_m)_{m \in \mathbb {N}}$ of degree $\mathrm {deg}(P_m) = d_m$ , such that

$$\begin{align*}b_k(\mathcal{X} \cap Z(P_m)) \ge a_{d_m},\end{align*}$$

where $b_k$ is the $k^{\text {th}}$ singular Betti number. This is a generalization of a result in [Reference Gwoździewicz, Kurdyka and Parusiński34]. Moreover, in their construction, which constructs a semianalytic $\mathcal {X}$ , $Z(P_m)$ intersects $\mathcal {X}$ transversely, thus making the topology of the intersection robust to small perturbations. See [Reference Natarajan49, Section 3.4] for a discussion of this issue.

1.1 Results: partitioning theorems

In this article, we make progress on the problem of obtaining an o-minimal polynomial partitioning theorem by proving partitioning theorems for the specific case of sets defined by Pfaffian functions (see Section 2.1 for precise definitions). Introduced by Khovanskiı̆ [Reference Khovanskiĭ39, Reference Khovanskiĭ40] for the study of fewnomials [Reference Khovanskiĭ41] and the second part of Hilbert’s sixteenth problem [Reference Ilyashenko35], Pfaffian functions satisfy triangular systems of first-order partial differential equations with polynomial coefficients. Pfaffian functions encompass a wide class of functions that contains, in particular, all Liouvillian functions (provided any sine or cosine functions involved are appropriately restricted). Pfaffian functions arise in many different mathematical areas. In model theory, they played an important role in Wilkie’s Theorem [Reference Wilkie67].

It is known that the Pfaffian structure, i.e., the smallest collection of sets containing all semi-Pfaffian sets and stable under all structure operations, is o-minimal. In fact, it is also known that Pfaffian functions over any o-minimal expansion of the real field form an o-minimal structure [Reference Speissegger61]. See [Reference Karpinski and Macintyre38] for an example in machine learning.

We generalize Theorem 1.1 by proving Theorem 3.2, which is a polynomial partitioning theorem for a set of semi-Pfaffian sets. Under a mild restriction on the Pfaffian chain, i.e., by restricting ourselves to algebraically independent Pfaffian chains (see Definition 2.6), we also prove Theorem 3.3, which is a partitioning theorem in which, instead of using a polynomial to partition the given set of semi-Pfaffian sets, we use another Pfaffian function to partition; henceforth, we shall refer to this as Pfaffian partitioning. Below, in Theorem 1.4, we only state instantiations of Theorem 3.2 and Theorem 3.3 because it becomes easier to contrast them with Theorem 1.1.

Notation 1.3 (Pfaffian functions)

We denote by $\mathrm {Pfaff}_n(\alpha ,\beta ,r)$ the set of Pfaffian functions in n variables with chain-degree $\alpha $ , degree $\beta $ , and order at most r. Also, if $\vec {q}$ is a Pfaffian chain of order r and chain-degree $\alpha $ , then $\mathrm {Pfaff}_{n}(\beta , \vec {q}) \subseteq \mathrm {Pfaff}_n(\alpha , \beta , r)$ denotes the set of all Pfaffian functions in $\mathrm {Pfaff}_n(\alpha ,\beta ,r)$ defined with respect to (w.r.t.) $\vec {q}$ . See Section 2.1 for the definition of Pfaffian functions and related concepts, including the notion of an algebraically independent Pfaffian chain.

Theorem 1.4 (Polynomial and Pfaffian partitioning of semi-Pfaffian sets - instantiations of Theorem 3.2 and Theorem 3.3)

Let $\Gamma $ be a collection of semi-Pfaffian sets in $\mathbb {R}^n$ of dimension k, where each $\mathcal {X} \in \Gamma $ is defined by at most m Pfaffian functions from $\mathrm {Pfaff}_n(\alpha ,\beta ,r)$ .

  1. (I) For any $D \ge 1$ , there is a nonzero polynomial $P \in \mathbb {R}[X_1, \ldots , X_n]$ of degree at most D, and a constant $C = C(n, m, \alpha , \beta , r)$ , such that each connected component of $\mathbb {R}^n \setminus Z(P)$ intersects at most $C\beta _1$ elements of $\Gamma $ , where

    $$\begin{align*}\beta_1 = \begin{cases} \frac{|\Gamma|}{D^{n - k - r}} & \text{if } k> 0 \\ \frac{|\Gamma|}{D^n} & \text{if } k = 0 \end{cases}. \end{align*}$$
  2. (II) Suppose that $\vec {q}$ is an algebraically independent Pfaffian chain (see Definition 2.6) of chain-degree $\alpha $ and order r, and suppose that all Pfaffian functions involved in defining the elements of $\Gamma $ are defined w.r.t. $\vec {q}$ . For any $D \ge 1$ , there is a nonzero Pfaffian function $P' \in \mathrm {Pfaff}_n(D, \vec {q})$ , and a constant $C = C(n, m, \alpha , \beta , r)$ , such that each connected component of $\mathbb {R}^n \setminus Z(P')$ intersects at most $C\beta _2$ elements of $\Gamma $ , where

    $$\begin{align*}\beta_2 = \begin{cases} \frac{|\Gamma|}{D^{n-k}} & \text{if } k> 0 \\ \frac{|\Gamma|}{D^{n+r}} & \text{if } k = 0 \end{cases}.\end{align*}$$

In our general theorems that are stated and proved in Section 3.2, we have multiple collections of semi-Pfaffian sets, and we establish the existence of a partitioning polynomial (Theorem 3.2), and a partitioning Pfaffian function (Theorem 3.3), that simultaneously partitions all collections. Polynomial partitioning of multiple collections of real-algebraic sets was considered in [Reference Blagojević, Blagojević and Ziegler13].

Given a set $\Gamma \subseteq \mathbb {R}^n$ , all partitioning theorems promise the existence of a function P such that $Z(P)$ partitions $\mathbb {R}^n$ ensuring that each connected component of $\mathbb {R}^n \setminus Z(P)$ intersects a small fraction of the elements of $\Gamma $ ; in what follows, the exact value of this fraction will be called the quantitative guarantee of the theorem.

Remark 1.5 (Polynomial partitioning vs Pfaffian partitioning)

Theorem 1.4-(I) generalizes Theorem 1.1 since we recover Theorem 1.1 by setting $r=0$ (c.f. Example 2.5-(I)). However, as the value of r gets closer to $n-k$ , the quantitative guarantee of Theorem 1.4-(I) approaches the trivial guarantee, i.e., $|\Gamma |$ . On the other hand, it is noteworthy that when $k \ge 1$ , the quantitative guarantee in Pfaffian partitioning of a set of semi-Pfaffian sets, i.e., Theorem 1.4-(II), matches the quantitative guarantee of polynomial partitioning of a set of semi-algebraic sets in Theorem 1.1, irrespective of the value of r; additionally, when $k=0$ , the quantitative guarantee is even better than what is obtained in Theorem 1.1. However, the caveat is that the object that accomplishes the partitioning of our set of semi-Pfaffian sets in Theorem 1.4-(II), namely the zero set of the Pfaffian function $P'$ , is topologically more complex than the real algebraic set that partitions the set of semi-algebraic sets in Theorem 1.1, which is the zero set of the polynomial P. Having said that, by using some auxiliary tools developed in Section 2.3 and due to the existence of Khovanskiı̆’s Bézout type bound for Pfaffian sets, we can control the complexity of the partitioning Pfaffian set by utilising constant-degree Pfaffian functions to partition our set of semi-Pfaffian sets. Pfaffian partitioning with constant degree Pfaffian functions is used in all applications shown in Section 4; see Section 4.3.2 for more details regarding the complexity of the partitioning set and how it influences our applications.

As discussed in Remark 1.2, the key step for proving Theorem 1.4 is to prove a semi-Pfaffian analogue of Theorem 2.14. In the statement of Theorem 1.6 below, $b_0$ denotes the $0^{\text {th}}$ Betti number, i.e., the number of connected components.

Theorem 1.6 (Semi-Pfaffian Barone-Basu type bound)

Let $\vec {q}$ be a Pfaffian chain of chain-degree $\alpha $ and order r. Let $\mathcal {X} \subseteq \mathbb {R}^n$ be a k-dimensional semi-Pfaffian set defined by at most m Pfaffian functions from $\mathrm {Pfaff}_n(\beta , \vec {q})$ . Then for any $P \in \mathrm {Pfaff}_n(D, \vec {q})$ , where $D \ge 1$ , there exists a constant $C = C(n, m, \alpha , \beta , r)$ such that

$$\begin{align*}b_0\left(\mathcal{X} \cap (\mathbb{R}^n \setminus Z(P))\right) \le CD^{k+r}.\end{align*}$$

Setting $r=0$ in Theorem 1.6 shows that it generalizes Theorem 2.14. The bound in Theorem 1.6, which obviously also works if $P \in \mathbb {R}[X_1, \ldots , X_n]$ of degree at most D, is obtained by setting up a system of Pfaffian equations such that the number of zeros of the system gives an upper bound on the number of connected components of $\mathcal {X} \cap (\mathbb {R}^n \setminus Z(P))$ ; the key to counting the number of solutions of this system is Khovanskiı̆’s Bézout type bound for Pfaffian sets, Theorem 2.11.

Remark 1.7 (Regarding the ‘r’ in the exponent of D in bound of Theorem 1.6)

In Khovanskiı̆’s Bézout type bound for Pfaffian functions (Theorem 2.11), the exponent in the bound depends on the order r of the Pfaffian chain, i.e., the largest degree of the Pfaffian functions involved is raised to the exponent r. It is this dependence that appears in the bound in Theorem 1.6. On the one hand, Khovanskiı̆’s bound is known to be at least exponential in r. On the other hand, while Theorem 2.14 is indeed asymptotically tight, we do not prove that Theorem 1.6 is asymptotically tight.

Remark 1.8 (Asymptotics in theorem statements)

In our main theorems, there are a number of different parameters involved: chain-degree $\alpha $ , degree $\beta $ , order r, the number of equations defining our Pfaffian sets m, cardinality of the set of varieties/Pfaffian sets $|\Gamma |$ , etc. These parameters are customarily separated into two categories – algebraic (e.g., degree $\beta $ , order r) and combinatorial (e.g., $|\Gamma |$ ). In general, one could be interested in various asymptotic regimes for the parameters involved. However, for several applications, especially in incidence combinatorics, discrete geometry, and harmonic analysis, the situation of relevance is where $|\Gamma |$ goes to infinity while the other parameters remain fixed. In other words, there is a huge number of objects, but the complexity of the individual objects is small. This motivates the treatment of certain parameters as part of the constants in our statements.

1.2 Results: applications of partitioning theorems

In Section 4, we derive two types of applications of our main theorems and other auxiliary tools proved in Section 2.3. In our first application, we study incidences between points and Pfaffian curves in $\mathbb {R}^n$ . Incidence problems are a major topic in discrete geometry that are interesting in their own right and also have links to other mathematical problems, such as the sum-product problem [Reference Elekes21] and the Kakeya conjectures [Reference Wolff68]; see [Reference Dvir20] for a survey of the area. In Theorem 4.1, we prove a Szemerédi-Trotter type theorem where, under nondegeneracy assumptions, we provide an upper bound on the number of incidences between points and Pfaffian curves in $\mathbb {R}^n$ . In Corollary 4.2, which is an instantiation of Theorem 4.1, we prove an upper bound on the number of incidences between points and Pfaffian curves in $\mathbb {R}^2$ . Corollary 4.2 has milder nondegeneracy assumptions, and, up to a small $\varepsilon $ factor, generalizes [Reference Sheffer58, Theorem 3.3] and the original Szemerédi-Trotter theorem [Reference Szemerédi and Trotter62], which is perhaps the first nontrivial incidence result.

The setting of Corollary 4.2 is not new in the literature, as both [Reference Basu and Raz9] and [Reference Chernikov, Galvin and Starchenko15] prove Szemerédi-Trotter type theorems in arbitrary o-minimal/distal structures, albeit with different formulations and different nondegeneracy assumptions. We would like to emphasize that the techniques in [Reference Basu and Raz9, Reference Chernikov, Galvin and Starchenko15] are quite different and arguably more involved; our partitioning theorem allows us to prove Pfaffian Szemerédi-Trotter type theorems by following the structure of the simpler methods used in the algebraic case.

As our second application, in Theorem 4.5, we derive the first upper bound on the number of joints formed by n sets of Pfaffian curves in $\mathbb {R}^n$ . The joints problem (see Definition 4.4 for a precise definition of joints) was originally posed in [Reference Chazelle, Edelsbrunner, Guibas, Pollack, Seidel, Sharir and Snoeyink14] and improved upon in multiple works; for example, [Reference Sharir56, Reference Feldman and Sharir23], with the final word on the number of joints for lines in $\mathbb {R}^3$ in [Reference Guth and Katz32]. The problem of counting joints is again of intrinsic mathematical value and is also motivated by applications in computational geometry related to the hidden surface removal problem; furthermore, connections between the joints problem and the Kakeya problem have been successfully exploited [Reference Wolff68, Reference Bennett, Carbery and Tao11].

Theorem 4.5 is a generalization of the results of [Reference Kaplan, Sharir and Shustin37, Reference Quilodrán53] and [Reference Yang69, Theorem 9]. As before, we are able to prove a generalization by following the structure of the proof in the algebraic case, albeit with appropriate modifications.

We discuss our applications and finer technical points regarding them in more detail in Section 4.3. Our applications, i.e., Theorem 4.1, Corollary 4.2, and Theorem 4.5, demonstrate the efficacy of our partitioning theorems and other auxiliary results in producing Pfaffian versions of discrete geometry theorems. The generality of our main theorems suggests that it is reasonable to expect that there will be many more applications to combinatorial problems involving semi-Pfaffian sets.

2 Preliminaries

Notation 2.1 (General notation)

$\mathbb {N}$ denotes the set of natural numbers $\{0, 1, 2, \ldots \}$ and $\mathbb {N}_{>0}$ the positive integers. We use the notation $[n] := \{1, \ldots , n\}$ . We denote by $\mathbb {R}[X_1, \ldots , X_n]_{(\le D)}$ the vector subspace of the polynomial ring $\mathbb {R}[X_1,\dots ,X_n]$ containing polynomials of degree at most D. We use the notation $\mathrm {x}=(x_1,\dots ,x_n)\in \mathbb {R}^n$ for a vector.

2.1 Basic definitions

Pfaffian functions are real or complex analytic functions that satisfy triangular systems of first-order partial differential equations with polynomial coefficients.

Definition 2.2 (Pfaffian functions)

Let $\mathcal {U} \subseteq \mathbb {R}^n$ be an open set. A Pfaffian chain of order $r \in \mathbb {N}$ and chain-degree $\alpha \in \mathbb {N}_{>0}$ over $\mathcal {U}$ is a sequence of functions $\vec {q} = (q_1, \ldots , q_r)$ , $q_i \in C^{\infty }(\mathcal {U})$ , such that there exist polynomials $P_{i, j} \in \mathbb {R}[X_1, \ldots , X_n, Y_1, \ldots , Y_j]_{(\le \alpha )}$ verifying

(2.1) $$ \begin{align} \,\mathrm{d}q_j(\mathrm{x}) = \sum_{i=1}^n P_{i,j}(\mathrm{x}, q_1(\mathrm{x}), \ldots, q_j(\mathrm{x}))\,\,\mathrm{d}x_i, \end{align} $$

for all $j \in [r]$ .

Let $\beta \in \mathbb {N}_{>0}$ . A function $f(\mathrm {x}) = P(\mathrm {x}, q_1(\mathrm {x}), \ldots , q_r(\mathrm {x}))$ , where $P \in \mathbb {R}[X_1, \ldots , X_n, Y_1, \ldots , Y_r]_{(\le \beta )}$ , is called a Pfaffian function. We will say that f is defined with respect to the chain $\vec {q}$ , has degree $\beta $ , has order r, and has chain-degree $\alpha $ . We will also sometimes say that f has degree $(\alpha , \beta )$ , and/or f has format $(\alpha , \beta , r)$ .

Remark 2.3. There are a few different definitions of Pfaffian functions in the literature. Definition 2.2, which matches the definition of Pfaffian functions in [Reference Gabrielov and Vorobjov25], is slightly more restrictive than the original definition in [Reference Khovanskiĭ41]. However, both definitions lead to nearly the same class of Pfaffian functions, albeit sometimes with different formats. In [Reference Miller and Speissegger47, Reference Lion and Rolin43], the authors allow for nonpolynomial coefficients in Equation (2.1).

Remark 2.4 (Domain $\mathcal {U}$ of Pfaffian functions [Reference Gabrielov and Vorobjov25])

The results in the paper are stated in terms of the format of the various semi-Pfaffian sets we consider. These results also depend on the domain $\mathcal {U}$ . While Definition 2.2 imposes no restrictions on $\mathcal {U}$ , thus allowing it to be arbitrarily complex causing the corresponding Pfaffian/semi-Pfaffian sets to also be arbitrarily complex, in the rest of the paper, we will always assume that the domain is ‘simple’, like $\mathbb {R}^n$ , $[0, 1]^n$ , $\left \{\mathrm {x} \boldsymbol {:} x_1> 0, \ldots , x_n > 0\right \}$ , or $\left \{\mathrm {x} \boldsymbol {:} \|\mathrm {x}\|_2^2 < 1\right \}$ . In more generality, we could have $\mathcal {U}$ be a semi-Pfaffian set defined by Pfaffian functions on a larger domain $\mathcal {U}' \supseteq \mathcal {U}$ , which in turn is defined by Pfaffian functions on some $\mathcal {U}" \supseteq \mathcal {U}'$ , and so on (see [Reference Gabrielov and Vorobjov25]).

Example 2.5 (Examples of Pfaffian functions)

  1. (I) A polynomial of degree D is a Pfaffian function w.r.t. the empty chain; format is $(\alpha , D, 0)$ for any integer $\alpha>0$ .

  2. (II) If $q(x) = e^{ax}$ , $\vec {q} = (q(x))$ is a Pfaffian chain over $\mathbb {R}$ given $\,\mathrm {d}q(x) = a q(x) \,\mathrm {d}x$ ; $\vec {q}$ has order $1$ and chain-degree $1$ . Any $P \in \mathbb {R}[X, e^{aX}]$ is Pfaffian function w.r.t. $\vec {q}$ .

  3. (III) As a generalization of the example above, for $i \in [r]$ , let $q_i(x) = e^{q_{i-1}(x)}$ , and $q_0(x) = ax$ . Since $\,\mathrm {d}q_i(x) = aq_1(x)\ldots q_i(x) \,\mathrm {d}x$ , $\vec {q} = (q_1, \ldots , q_r)$ is a Pfaffian chain of order r and chain-degree r. Consequently, any $P \in \mathbb {R}[X, e^{aX}, e^{e^{aX}}, \ldots ]$ is a Pfaffian function w.r.t. $\vec {q}$ .

  4. (IV) $\vec {q} = (\tan (x))$ is a Pfaffian chain of order $1$ and chain-degree $2$ in the domain , given $\,\mathrm {d}\tan (x) = \left (1 + \tan (x)^2\right ) \,\mathrm {d}x$ ; any $P \in \mathbb {R}[X, \tan (X)]$ is a Pfaffian function w.r.t. $\vec {q}$ .

  5. (V) $\vec {q} = \left (\frac {1}{x}, \ln (x)\right )$ is a Pfaffian chain on the domain $\mathbb {R}_{>0}$ since

    $$\begin{align*}\,\mathrm{d}\frac{1}{x} = -\left(\frac{1}{x}\right)^2 \,\mathrm{d}x, \qquad \text{and} \qquad \,\mathrm{d}\ln(x) = \frac{1}{x} \,\mathrm{d}x.\end{align*}$$
    Consequently, any $P \in \mathbb {R}\left [X, \frac {1}{X}, \ln (X)\right ]$ is a Pfaffian function.
  6. (VI) $\vec {q} = \left (\frac {1}{x}, x^m\right )$ for any $m \in \mathbb {R}$ is a Pfaffian chain since

    $$\begin{align*}\,\mathrm{d}\frac{1}{x} = -\left(\frac{1}{x}\right)^2 \,\mathrm{d}x, \qquad \text{and} \qquad \,\mathrm{d}x^m = m \cdot \frac{1}{x} \cdot x^m \,\mathrm{d}x. \end{align*}$$
    Thus any $P \in \mathbb {R}\left [X, \frac {1}{X}, X^m\right ]$ is a Pfaffian function w.r.t. $\vec {q}$ . Note that the domain is $\mathbb {R}_{>0}$ unless m is an integer.
  7. (VII) For any monomial $m_{i_1, \ldots , i_n} := a_{i_1, \ldots , i_n}x_1^{i_1}\ldots x_n^{i_n}$ with $a_{i_1, \ldots , i_n} \neq 0$ , $\left (\frac {1}{x_1}, \ldots , \frac {1}{x_n}, m_{i_1, \ldots , i_n}\right )$ is a Pfaffian chain on the domain $\{\mathrm {x} \in \mathbb {R}^n \boldsymbol {:} x_1\dots x_n \neq 0\}$ of order $n+1$ , and chain-degree $2$ , given that

    $$ \begin{align*} \,\mathrm{d}m_{i_1, \ldots, i_n} = m_{i_1, \ldots, i_n}\cdot \sum_{j=1}^n i_j \frac{1}{x_j} \,\mathrm{d}x_j, \qquad \text{and} \qquad \,\mathrm{d}\frac{1}{x_j} = -\left(\frac{1}{x_j}\right)^2 \,\mathrm{d}x_j. \end{align*} $$
    From this we can deduce that if f is a polynomial that is a sum of s such monomials, called a fewnomial of sparsity s, then f is a Pfaffian function (on the same domain as before) of order $n+s$ and degree $(2, 1)$ .

We define the following restricted class of Pfaffian chains.

Definition 2.6 (Algebraically independent Pfaffian chain)

A Pfaffian chain $\vec {q} = (q_1, \ldots , q_r)$ is called algebraically independent if, for all nonzero $P \in \mathbb {R}[Y_1, \ldots , Y_r]$ , $P(q_1(\mathrm {x}), \ldots , q_r(\mathrm {x}))$ is not identically zero.

Remark 2.7 (Regarding algebraic independence of Pfaffian chains)

We prove two types of partitioning theorems, namely partitioning using a polynomial and partitioning using a Pfaffian function. We place the requirement that the common Pfaffian chain be algebraically independent only in the latter case. This is solely to ensure that the partitioning Pfaffian function obtained is nonzero; the zero function trivially satisfies the promised conditions of the partitioning theorems but is worthless from the point of view of applications.

How restrictive is the condition of algebraic independence? All examples in Example 2.5, except (VI) and (VII), have algebraically independent Pfaffian chains. In fact, if in Example 2.5-(VI), $m \in \mathbb {R} \setminus \mathbb {Q}$ , then $\vec {q}$ would be algebraically independent (see below for a proof of this). Also, if the different elements of the Pfaffian chain contain different elementary functions, the Pfaffian chain will be algebraically independent.

Proof of algebraic independence of the Pfaffian chain in Example 2.5-(VI) when $m \in \mathbb {R} \setminus \mathbb {Q}$

We wish to prove that $\vec {q} = \left (\frac {1}{x}, x^m\right )$ is an algebraically independent Pfaffian chain whenever m is irrational.

Suppose, by way of contradiction, that for some fixed $m \in \mathbb {R} \setminus \mathbb {Q}$ , there is a polynomial $P \in \mathbb {R}[Y_1, Y_2]$ , of degree D, such that $P \neq 0$ and $P\left (\frac {1}{X}, X^m\right )$ is identically $0$ . Suppose that P is made of monomials $\{Y_1^{a_i}Y_2^{b_i}\}$ , with $a_i, b_i \in \mathbb {N}$ and $a_i + b_i \le D$ . On substitution, the monomial $Y_1^{a_i}Y_2^{b_i}$ becomes $X^{mb_i - a_i}$ . Notice that the map

$$\begin{align*}Y_1^{a_i}Y_2^{b_i} \mapsto X^{mb_i - a_i} \end{align*}$$

is injective; because if not, it would mean that there exist $a, b, a', b' \in \mathbb {N}$ such that $(a, b) \neq (a', b')$ and $mb' - a' = mb - a$ . This, in turn, would mean that $m = \frac {a' - a}{b' - b}$ , which is a contradiction to the assumption that $m \in \mathbb {R} \setminus \mathbb {Q}$ . Consequently, if $P\left (\frac {1}{X}, X^m\right )$ is identically $0$ , it means that $mb_i - a_i = 0$ for all $(a_i, b_i)$ ; which is also a contradiction to the assumption that $m \in \mathbb {R} \setminus \mathbb {Q}$ .

Definition 2.8 (Pfaffian sets)

A Pfaffian set in an open domain $\mathcal {U} \subseteq \mathbb {R}^n$ is a set of the form

$$\begin{align*}Z(g_1, \ldots, g_k) := \{\mathrm{x} \in \mathcal{U} \boldsymbol{:} g_1(\mathrm{x}) = \cdots = g_k(\mathrm{x}) = 0\},\end{align*}$$

where $g_i: \mathbb {R}^n \rightarrow \mathbb {R}$ are Pfaffian functions.

Definition 2.9 (Semi-Pfaffian sets)

A set $\mathcal {X} \subseteq \mathbb {R}^n$ is called semi-Pfaffian in an open domain $\mathcal {U} \subseteq \mathbb {R}^n$ if it is the locus in $\mathcal {U}$ of a Boolean combination of equations $f = 0$ and inequalities $g> 0$ , where $f, g$ are Pfaffian functions having a common Pfaffian chain defined on $\mathcal {U}$ .

A basic semi-Pfaffian set is one where the corresponding Boolean combination is merely a conjunction of equations and strict inequalities. That is, a basic semi-Pfaffian set is of the form

$$ \begin{align*} \mathcal{X} = \{\mathrm{x} \in \mathcal{U} \colon g_1(\mathrm{x})=\cdots =g_s(\mathrm{x})=0, \ h_1(\mathrm{x})>0,\dots,h_t(\mathrm{x})>0\}, \end{align*} $$

where the $g_i$ and $h_j$ are Pfaffian functions having a common Pfaffian chain defined on $\mathcal {U}$ .

The complexity of a semi-Pfaffian set $\mathcal {X}$ is defined as the coordinatewise maximum of the formats of the Pfaffian functions that define $\mathcal {X}$ .

We say a semi-Pfaffian set has order r if it can be defined by Pfaffian functions defined w.r.t. the same Pfaffian chain of order not exceeding r.

Our proofs depend on the fact that semi-Pfaffian sets can be decomposed into semi-Pfaffian subsets, in such a way that the complexity of these subsets can be controlled. Relevant definitions and results below are based on [Reference Gabrielov and Vorobjov25], and we refer to that source for more background, proofs, and references.

Definition 2.10 (Weak stratification)

A weak stratification of a semi-Pfaffian set $\mathcal {X}$ is a partition of $\mathcal {X}$ into a disjoint union of smooth, not necessarily connected, possibly empty semi-Pfaffian subsets $\mathcal {X}_i$ , called strata. A stratification is basic if all strata are basic semi-Pfaffian sets that are effectively nonsingular, i.e., the system of equations and strict inequalities for each stratum $\mathcal {X}_i$ of codimension k includes a set of k Pfaffian functions $h_{1}, \ldots , h_{k}$ vanishing on $\mathcal {X}_i$ , such that their gradients are linearly independent at every point in $\mathcal {X}_i$ .

2.2 Pfaffian and real algebraic geometry preliminaries

In this section, we collect a few well-known results in Pfaffian and real-algebraic geometry that we will use in the rest of the paper.

An important result in the theory of Pfaffian functions is Khovanskiı̆’s Bézout type theorem for Pfaffian sets.

Theorem 2.11 (Khovanskiı̆’s theorem [Reference Khovanskiĭ41, §3.12, Corollary 5])

Let $\vec {q}$ be a Pfaffian chain of order r and chain-degree $\alpha $ , where the functions in the Pfaffian chain depend only on $\xi \le n$ variables. Let $f_1, \ldots , f_n$ be Pfaffian functions on an open set $\mathcal {U} \subseteq \mathbb {R}^n$ , where $f_i \in \mathrm {Pfaff}_n(\beta _i, \vec {q})$ . The number of nondegenerate solutions of the system $\{\mathrm {x} \in \mathcal {U} \boldsymbol {:} f_1(\mathrm {x}) = \cdots = f_n(\mathrm {x}) = 0\}$ is bounded from above by

$$\begin{align*}2^{\binom{r}{2}}\beta_1 \cdots \beta_n\left(\min\{\xi, r\}\alpha + \beta_1 + \cdots + \beta_n - n + 1\right)^r.\end{align*}$$

From hereon, we will omit reference to the domain $\mathcal {U}$ and implicitly assume that all occurring Pfaffian functions are defined on the same domain.

The following proposition establishes the existence of a basic weak stratification for any semi-Pfaffian set.

Proposition 2.12 ([Reference Gabrielov and Vorobjov25]*Theorem 6.2, Smooth stratification)

Let $\vec {q}$ be a Pfaffian chain of order r and chain-degree $\alpha $ . Any semi-Pfaffian set $\mathcal {X} \subset \mathbb {R}^n$ defined by m Pfaffian functions in $\mathrm {Pfaff}_n(\beta , \vec {q})$ has a finite basic weak stratification. There exist constants $C_1=C_1(\alpha , r, m, n)$ and $C_2=C_2(r,n)$ such that the number of connected components of all strata and their complexities is bounded from above by $C_1\beta ^{C_2}$ .

Next, we state a couple of well-known results in real algebraic geometry. The first is a bound on the total number of connected components in $\mathbb {R}^n$ of the complement of the zero set of a polynomial $P \in \mathbb {R}[X_1, \ldots , X_n]$ .

Theorem 2.13 [Reference Petrovskii and Oleinik52, Reference Milnor48, Reference Thom63]

For any $P \in \mathbb {R}[X_1, \ldots , X_n]$ of degree D, there exists a constant $C = C(n)$ such that

$$\begin{align*}b_0(\mathbb{R}^n \setminus Z(P)) \le CD^n.\end{align*}$$

From Theorem 2.13, we have that the zero set of any $P \in \mathbb {R}[X_1, \ldots , X_n]$ breaks $\mathbb {R}^n$ into at most $\mathcal {O}_n(D^n)$ connected components. For $k < n$ , given a k-dimensional real-algebraic set $\mathcal {X}$ , the following theorem gives an upper bound on the number of connected components of $\mathbb {R}^n \setminus Z(P)$ that $\mathcal {X}$ intersects. We state the formulation of [Reference Solymosi and Tao60, Theorem A.2] which is implied by a more general result [Reference Barone and Basu6, Theorem 4].

Theorem 2.14 ([Reference Barone and Basu6, Theorem 4], [Reference Solymosi and Tao60, Theorem A.2])

Suppose that $\mathcal {X} \subseteq \mathbb {R}^n$ is a k-dimensional real-algebraic set defined by m polynomial equations, each of degree at most d. Given any polynomial $P \in \mathbb {R}[X_1, \ldots , X_n]$ of degree at most D, then there is a constant $C = C(n, m, d)$ such that

$$\begin{align*}b_0\left(\mathcal{X} \cap (\mathbb{R}^n \setminus Z(P))\right) \le CD^k.\end{align*}$$

The key to our results is a generalization of the above bound to semi-Pfaffian sets.

2.3 Auxiliary tools

In this section, we prove a few auxiliary results that we shall use in proving our main theorems, as well as applications of those theorems.

Since Theorem 2.11 only gives a bound on the number of nondegenerate solutions of a Pfaffian system, we need the following corollary.

Corollary 2.15 (of Theorem 2.11)

Let $\vec {q}$ be a Pfaffian chain of order r and chain-degree $\alpha $ , where the functions in the Pfaffian chain depend only on $\xi \le n$ variables. Let $f_1, \ldots , f_n$ be Pfaffian functions on an open set $\mathcal {U} \subseteq \mathbb {R}^n$ , where $f_i \in \mathrm {Pfaff}_n(\beta _i, \vec {q})$ . Suppose we have that for all $i \in [n]$ , $\mathrm {dim}\, Z(f_i) = n-1$ and $\mathrm {dim}\, Z(f_1, \ldots , f_i) = n - i$ . Then the number of solutions of the system $M:= \{\mathrm {x} \in \mathcal {U} \boldsymbol {:} f_1(\mathrm {x}) = \cdots = f_n(\mathrm {x}) = 0\}$ is bounded from above by

$$\begin{align*}2^{n + \binom{r+1}{2}}\beta_1 \cdots \beta_n\left(\min\{\xi, r\}\alpha + \beta_1 + \cdots + \beta_n - n + 1\right)^r.\end{align*}$$

Proof of Corollary 2.15

First we introduce the following definition from [Reference Gabrielov and Vorobjov26, Definition 1.7]. Let ${\mathcal P}= {\mathcal P}(\varepsilon _1, \ldots ,\varepsilon _\ell )$ be a predicate (property) over $(0,1)^\ell $ . We say that $\mathcal P$ holds for $0 < \varepsilon _\ell \ll \varepsilon _{\ell -1} \ll \cdots \ll \varepsilon _1 \ll 1$ if there exist functions $t_k:\> (0,1)^{k-1} \to (0,1)$ , definable in real analytic structure $\mathbb {R}_{an}$ (see, e.g., [Reference Binyamini and Novikov12, Section 2.1.2]), where $k=1, \ldots , \ell $ (with $t_1$ being a positive constant) such that $\mathcal P$ holds for any sequence $\varepsilon _1, \ldots ,\varepsilon _\ell $ satisfying $0< \varepsilon _k <t_k(\varepsilon _{k-1}, \ldots ,\varepsilon _1)$ .

For real numbers $\varepsilon _1, \ldots , \varepsilon _n$ consider the set $Z(f_1^2- \varepsilon _1, \ldots , f_n^2-\varepsilon _n)$ . We prove, by induction on $\ell $ , that for $0< \varepsilon _\ell \ll \cdots \ll \varepsilon _1 \ll 1$ all solutions x in $Z(f_1^2- \varepsilon _1, \ldots , f_\ell ^2-\varepsilon _\ell )$ are regular, i.e., the matrix $[\nabla f_1^2 ({\mathrm {x}}), \ldots ,\nabla f_\ell ^2 ({\mathrm {x}})]$ has full rank, and tend to solutions in $Z(f_1, \ldots , f_\ell )$ as $\varepsilon _i \to 0$ .

The base case $\ell =1$ is obvious. For $\ell < n$ , assume that all points in $Z(f_1^2-\varepsilon _1, \ldots ,f_\ell ^2-\varepsilon _{\ell })$ are regular. Consider the restriction $g_{\ell +1}$ of $f_{\ell +1}^2$ to $Z(f_1^2-\varepsilon _1, \ldots ,f_\ell ^2-\varepsilon _{\ell })$ . Note, that $Z(f_1^2-\varepsilon _1, \ldots ,f_\ell ^2-\varepsilon _{\ell }, g_{\ell +1})$ is $(n-\ell -1)$ -dimensional and the function $g_{\ell +1}$ attains the minimum ( $ = 0$ ) on $Z(f_1^2-\varepsilon _1, \ldots ,f_\ell ^2-\varepsilon _{\ell })$ at its critical points. Thus, $g_{\ell +1}-\varepsilon _{\ell +1}$ , for small $\varepsilon _{\ell +1}>0$ , does not have critical points on $Z(f_1^2-\varepsilon _1, \ldots ,f_\ell ^2-\varepsilon _{\ell })$ with value $0$ , hence all points in $Z(f_1^2-\varepsilon _1, \ldots ,f_\ell ^2-\varepsilon _{\ell }, f_{\ell +1}^2-\varepsilon _{\ell + 1})$ are regular.

Finally, applying Theorem 2.11 to $Z(f_1^2-\varepsilon _1, \ldots , f_n^2- \varepsilon _n)$ , noticing that the degrees of $f_i^2 - \varepsilon _i$ are $2\beta _i$ , completes the proof.

Notation 2.16. For any semi-Pfaffian set $\mathcal {X} \subset \mathbb {R}^n$ , let $\mathcal {X}^{\text {smooth}}$ be the subset of all points $x \in \mathcal {X}$ such that the neighborhood of x in $\mathcal {X}$ is a smooth manifold.

Corollary 2.17 and Corollary 2.18 below use the stratification result in Proposition 2.12 to decompose a Pfaffian set into smooth sets and also provide some quantitative results on the topology of Pfaffian sets.

Corollary 2.17 (of Proposition 2.12)

Let $\vec {q}$ be a Pfaffian chain of chain-degree $\alpha $ and order r. Let $\mathcal {X} \subseteq \mathbb {R}^n$ be a k-dimensional Pfaffian set defined by at most m Pfaffian functions from $\mathrm {Pfaff}_n(\beta , \vec {q})$ . Then there exists a constant $C = C(n, m, \alpha , \beta , r)$ such that one can cover $\mathcal {X}$ as

$$\begin{align*}\mathcal{X} = \bigcup_{i=1}^{C} \mathcal{X}_i^{\text{smooth}},\end{align*}$$

where each $\mathcal {X}_i \subseteq \mathcal {X}$ is a Pfaffian set of dimension at most k defined by at most C Pfaffian functions from $\mathrm {Pfaff}_n(C, \vec {q})$ .

Proof. Using Proposition 2.12, perform a basic weak stratification of $\mathcal {X}$ to obtain strata $S_1, \ldots , S_C$ . Each stratum $S_i$ is represented as a set of points in $\mathbb {R}^n$ satisfying a system of equations $E_{i}$ , and strict inequalities $I_{i}$ . Define $\mathcal {X}_i$ as the intersection of the set defined by $E_{i}$ and $\mathcal {X}$ . Then we claim that $\mathcal {X}_i$ are Pfaffian sets, and that the $\mathcal {X}_i^{\text {smooth}}$ form the required covering of $\mathcal {X}$ .

Indeed, for all $i \in [C]$ , $\mathcal {X}_i \subseteq \mathcal {X}$ , so $\bigcup _{i=1}^C X_i^{\text {smooth}} \subseteq \mathcal {X}$ . Also, for any $x \in \mathcal {X}$ , there exists $S_i$ such that $x \in S_i$ . Since $S_i \subseteq \mathcal {X}_i^{\text {smooth}} \subseteq \mathcal {X}$ , we have that $x \in \mathcal {X}_i^{\text {smooth}} \subseteq \bigcup _{i=1}^C \mathcal {X}_i^{\text {smooth}}$ , proving that $\mathcal {X} \subseteq \bigcup _{i=1}^C \mathcal {X}_i^{\text {smooth}}$ . Consequently, we have

$$\begin{align*}\mathcal{X} = \bigcup_{i=1}^C \mathcal{X}_i^{\text{smooth}},\end{align*}$$

as required.

Also, it is true that $\mathrm {dim} \, \mathcal {X}_i \le k$ , and the conditions required on the complexity of $\mathcal {X}_i$ follow from the guarantee of Proposition 2.12.

Corollary 2.18 (of Proposition 2.12)

Let $\vec {q}$ be a Pfaffian chain of chain-degree $\alpha $ and order r, and let $P \in \mathrm {Pfaff}_n(\beta , \vec {q})$ . There exist constants $C_1=C_1(\alpha , \beta , r, n)$ and $C_2=C_2(r,n)$ such that the number of irreducible (over $\mathbb {R}$ ) components of $Z(P)$ is at most $C_1 \beta ^{C_2}$ .

Proof. Applying Proposition 2.12 to $Z(P)$ , we conclude that the number of connected components of smooth strata in the stratification of $Z(P)$ does not exceed $C_1 \beta ^{C_2}$ , where both $C_1$ and $C_2$ are the constants coming from Proposition 2.12 itself.

Let $Z(Q)$ be an irreducible component of $Z(P)$ having dimension d. We have that there is a connected component Y of a d-dimensional stratum such that $\dim (Z(Q) \cap Y) = d$ , i.e., the function Q vanishes on a d-dimensional disc in Y. By the identity theorem (see for instance [Reference Krantz and Parks42, Section 1.2]), Q vanishes on all of Y, hence $Y \subseteq Z(Q)$ . Thus, it follows that the number of irreducible components of $Z(P)$ is at most the number of connected components of smooth strata of $Z(P)$ . Consequently, $C_1\beta ^{C_2}$ is an upper bound on the number of irreducible components.

3 Polynomial and Pfaffian partitioning of a set of semi-Pfaffian sets

In this section, we prove our main theorem, a partitioning theorem for a set of semi-Pfaffian sets (Theorem 1.4).

3.1 Smooth decomposition of semi-Pfaffian sets

The first step towards Theorem 1.4 is to establish an upper bound on the number of connected components of $\mathcal {X} \cap (\mathbb {R}^n \setminus Z(P))$ for a semi-Pfaffian set $\mathcal {X}$ and a polynomial P (Theorem 1.6). The following lemma describes a procedure to obtain a smooth decomposition of semi-Pfaffian sets using the weak smooth stratification result stated in Proposition 2.12.

Lemma 3.1. Let $\mu>0$ and let $\mathcal {X} \subseteq (-\mu , \mu )^n \subseteq \mathbb {R}^n$ be a semi-Pfaffian set of order r. Then there exists a partition of $\mathcal {X}$ into semi-Pfaffian subsets, called cells, such that each n-dimensional cell is an open set in $(-\mu , \mu )^n$ , while each k-dimensional set for $k < n$ is the graph of a smooth map $\mathbf {f}=(f_1, \ldots ,f_{n-k})$ defined on an open set in one of the k-dimensional faces of $(-\mu , \mu )^n$ . Moreover, each cell has order r, and is effectively nonsingular (see Definition 2.10). The number of cells, the number of functions defining cells (open sets and graphs), and the complexities of cells are bounded from above by an explicit function of the complexity of $\mathcal {X}$ .

Proof. The proof is by induction on $\dim \mathcal {X}=k$ . If $k=0$ , then $\mathcal {X}$ consists of a finite number of points, and each of these constitutes a cell that satisfies the required properties (the bound on the number of points is provided by Theorem 2.11). Assume $k>0$ . By Proposition 2.12, there exists a weak, smooth stratification of $\mathcal {X}$ into a finite number of basic semi-Pfaffian sets. Let $\mathcal {X}_{<k}$ be the union of all strata of dimension less than k.

Consider a stratum $\mathcal {C}$ with $\dim \mathcal {C}=k$ . We first show that we can subdivide $\mathcal {C}$ into strata that can be represented as graphs, and later argue that the complexity of the functions defining the graphs can be controlled in terms of the parameters defining the semi-Pfaffian set. Since the stratification is basic, there exist Pfaffian functions $h_1,\dots ,h_{n-k}$ such that $h_i|_{\mathcal {C}}=0$ and $\mathrm {d}h_1\wedge \cdots \wedge \mathrm {d}h_{n-k}\neq 0$ on $\mathcal {C}$ . For $\{ i_1, \ldots , i_m \} \subset [n]$ , define the minor

$$ \begin{align*} \mathbf{J}(i_1, \ldots, i_m ):= \left( \begin{array}{ccc} \partial h_1/\partial x_{i_1} & \cdots & \partial h_1 / \partial x_{i_m}\\ \vdots & \ddots & \vdots \\ \partial h_{n-k}/ \partial x_{i_1} & \cdots & \partial h_{n-k}/ \partial x_{i_m} \end{array} \right). \end{align*} $$

Note that $\mathbf {J}(1,2, \ldots , n )$ has rank $n-k$ on $\mathcal {C}$ .

Fix a total order (for example, lexicographic) on the family of all subsets of $\{1, 2, \ldots ,n \}$ with a cardinality of $n-k$ . Suppose, for definiteness, that the minimal set, with respect to this order, is $\{ 1, 2, \ldots , n-k\}$ .

Define

$$ \begin{align*} \mathcal{C}_{1, \ldots, n-k} := \mathcal{C} \cap \{ \det \mathbf{J}(1, \ldots, n-k)=0 \}. \end{align*} $$

Observe that $\mathcal {C} \setminus \mathcal {C}_{1, \ldots , n-k}$ is either empty or k-dimensional, while $\dim \mathcal {C}_{1, \ldots , n-k} \le k$ is possibly empty. If $\dim \mathcal {C}_{1, \ldots , n-k} < k$ , then reset $\mathcal {X}_{<k}:= \mathcal {X}_{<k} \cup \mathcal {C}_{1, \ldots , n-k}$ . Otherwise, perform a weak stratification of $\mathcal {C}_{1, \ldots , n-k}$ , reset $\mathcal {X}_{<k}$ by taking the union of the current $\mathcal {X}_{<k}$ with strata of dimension less than k, and denote by $\widehat {\mathcal {C}}_{1, \ldots , n-k}$ the k-dimensional stratum.

Assume that we constructed the smooth k-dimensional $\widehat {\mathcal {C}}_{j_1, \ldots , j_{n-k}}$ for a set $\{ j_1, \ldots ,j_{n-k} \}$ in the chosen ordering of subsets of $\{1, 2, \ldots ,n \}$ . Let $\{ i_1, \ldots i_{n-k} \}$ be the next set in the ordering. Define

$$ \begin{align*} \mathcal{C}_{i_1, \ldots, i_{n-k}} := \widehat{\mathcal{C}}_{j_1, \ldots, j_{n-k}} \cap \{ \det \mathbf{J}(i_1, \ldots, i_{n-k})=0 \}. \end{align*} $$

Observe that $\widehat {\mathcal {C}}_{j_1, \ldots , j_{n-k}} \setminus \mathcal {C}_{i_1, \ldots , i_{n-k}}$ is either empty or k-dimensional, while $\dim \mathcal {C}_{i_1, \ldots , i_{n-k}} \le k$ is possibly empty.

If $\dim \mathcal {C}_{i_1, \ldots , i_{n-k}} < k$ , then reset $\mathcal {X}_{<k}:= \mathcal {X}_{<k} \cup \mathcal {C}_{i_1, \ldots , i_{n-k}}$ . Otherwise, perform a weak stratification of $\mathcal {C}_{i_1, \ldots , i_{n-k}}$ , reset $\mathcal {X}_{<k}$ by taking the union of the current $\mathcal {X}_{<k}$ with strata of dimension less than k, and denote by $\widehat {\mathcal {C}}_{i_1, \ldots , i_{n-k}}$ the k-dimensional stratum.

Continue this process by passing from one subset in the chosen ordering of subsets of $\{1, 2, \ldots ,n \}$ to the next. The process will terminate if either the current $\mathcal {C}_{i_1, \ldots , i_{n-k}}$ has dimension less than k, or is empty. The latter case occurs when for the previous set $\{ j_1, \ldots , j_{n-k} \}$ the stratum $\widehat {\mathcal {C}}_{j_1, \ldots , j_{n-k}}$ does not have points with vanishing $\det \mathbf {J}(i_1, \ldots , i_{n-k})$ . The process always terminates since effective nonsingularity of $\mathcal {C}$ implies that the rank of the $(n-k) \times n$ - matrix $\mathbf {J}(1, \ldots n)$ is maximal at every point of $\mathcal {C}$ .

By the implicit function theorem, each nonempty set $\widehat {\mathcal {C}}_{j_1, \ldots , j_{n-k}} \setminus \mathcal {C}_{i_1, \ldots , i_{n-k}}$ is a disjoint union of graphs of smooth maps defined on their projections to the k-dimensional face of $(-\mu , \mu )^n$ in the subspace of coordinates $x_{\ell _1}, \ldots , x_{\ell _k}$ , where

$$\begin{align*}\{ \ell_1, \ldots , \ell_k \}= \{ 1, \ldots n \} \setminus \{ i_1, \ldots , i_{n-k} \}.\end{align*}$$

Hence, in the terminating ordered sequence of subsets of $\{1, 2, \ldots ,n \}$ the sets

$$ \begin{align*} (\mathcal{C} \setminus \mathcal{C}_{1, \ldots, n-k}), \ldots , (\widehat{\mathcal{C}}_{j_1, \ldots, j_{n-k}} \setminus \mathcal{C}_{i_1, \ldots, i_{n-k}}) \end{align*} $$

are disjoint unions of k-dimensional graphs of smooth maps, while the complement of their union in $\mathcal {C}$ has a dimension less than k. Applying the inductive hypothesis to $\mathcal {X}_{<k}$ , we obtain a partition of $\mathcal {X}$ into graphs of smooth maps.

In order to see that the functions $f_i$ provided by the implicit function theorem are Pfaffian functions whose parameters can be bounded in terms of the equations $h_1,\dots ,h_k$ defining the original strata, note that from the implicit function theorem we can extract an expression for the partial derivatives $\partial f_i/\partial x_j$ in terms of the partial derivatives of the $h_i$ . Complexities of functions $h_i$ are bounded from above by Proposition 2.12.

3.2 Proofs of Theorem 1.6 and Theorem 1.4

We now have all the ingredients required to prove Theorem 1.6 and Theorem 1.4.

Proof of Theorem 1.6

For all sufficiently large integers $\mu>0$ , the number of connected components of $\mathcal {X} \cap (\mathbb {R}^n \setminus Z(P))$ does not exceed (in fact, coincides with) the number of connected components of $\mathcal {X} \cap (-\mu , \mu )^n \cap (\mathbb {R}^n \setminus Z(P))$ . Therefore, without loss of generality, we will assume in this proof that $\mathcal {X} \subseteq (-\mu , \mu )^n$ for some $\mu>0$ . Apply Lemma 3.1 to get a partition $\mathcal {C} = \{\mathcal {C}_1, \ldots , \mathcal {C}_v\}$ of $\mathcal {X}$ . We will establish that there is a constant $C' = C'(n, k, m, \alpha , \beta , r)$ such that, for any $\mathcal {C}_i \in \mathcal {C}$ ,

(3.1) $$ \begin{align} b_0(\mathcal{C}_i \setminus Z(P)) \le C'D^{k},\end{align} $$

completing the proof because, by Lemma 3.1, we have that $|\mathcal {C}|$ does not grow with D.

We proceed by induction on k. The base case of $k = 0$ is trivial, so suppose that $k \ge 1$ , and that (3.1) is proved for all $\mathcal {X}$ of dimension strictly less than k. Fix a k-dimensional smooth cell $\mathcal {C}_i$ . For this specific $\mathcal {C}_i$ , let $\alpha _1, \ldots , \alpha _{t}$ be the connected components of $\{\mathrm {x} \in \mathcal {C}_i \boldsymbol {:} P(\mathrm {x}) \neq 0\}$ . Obviously, this means that

(3.2) $$ \begin{align} b_0(\mathcal{C}_i \setminus Z(P)) \le t.\end{align} $$

We might as well assume that all $\alpha _i$ do not intersect the boundary $\partial \mathcal {C}_i$ , because the number of components that intersect $\partial \mathcal {C}_i$ is at most $\mathcal {O}_{n, \partial \mathcal {C}_i}(D^{k-1})$ by the induction hypothesis. Naturally, by the continuity of P, on any particular $\alpha _j$ , we will either have $\forall \mathrm {x} \in \alpha _j$ , $P(\mathrm {x})> 0$ , or $\forall \mathrm {x} \in \alpha _j$ , $P(\mathrm {x}) < 0$ . Consequently, P is guaranteed to have at least one critical point on each $\alpha _j$ .

Define $A := \{i \in [t] \boldsymbol {:} \forall \mathrm {x} \in \alpha _i, P(\mathrm {x})> 0 \}$ . Choose $\varepsilon> 0$ to be a regular value of P that is strictly smaller than the minimum of all the critical values of P on all $\alpha _j$ , $j \in A$ , which are all obviously positive because of how A is defined. Since $\varepsilon $ is positive and is also strictly smaller than the minimum critical value of P, for all $j \in A$ , $Z(P - \varepsilon ) \cap \alpha _j \neq \emptyset $ . Also, since $\varepsilon $ is a regular value of P, by the implicit function theorem, $Z(P - \varepsilon )$ is smooth. Given that by Lemma 3.1, all $\alpha _j$ , $j \in [t]$ , are also smooth, we can conclude that $Z(P - \varepsilon ) \cap \alpha _j$ is smooth. $Z(P - \varepsilon ) \cap \alpha _j$ is also guaranteed to be compact because $Z(P - \varepsilon )$ cannot touch $\partial \alpha _j$ for all $j \in A$ . Our goal will be to find an upper bound on $|A|$ . Notice that $[t] \setminus A$ is exactly the set $\{i \in [t] \boldsymbol {:} \forall \mathrm {x} \in \alpha _i, P(\mathrm {x}) < 0 \}$ . Thus by analyzing similarly as what is above with $\varepsilon < 0$ , and looking at $P + \varepsilon $ instead of $P - \varepsilon $ , we will be able to obtain an upper bound on the cardinality of the complement of A in $[t]$ ; summing up both upper bounds will give an upper bound on t.

By the definition of the partition in Lemma 3.1, $\mathcal {C}_i$ is the graph of a smooth map $\mathbf {f}_i=(f_{i,1}, \ldots f_{i, n-k})$ defined on the projection of $\mathcal {C}_i$ on a k-dimensional face F of $(-\mu , \mu )^n$ . Assume, for definiteness, that F has coordinates $x_1, \ldots , x_k$ . Then, $Z(P - \varepsilon ) \cap \mathcal {C}_i$ is $(k-1)$ -dimensional smooth Pfaffian set, and its projection $\mathrm {proj}\ (Z(P - \varepsilon ) \cap \mathcal {C}_i)$ on F is a smooth embedding. In particular, $\mathrm {proj}\ (Z(P - \varepsilon ) \cap \mathcal {C}_i)$ is a smooth compact manifold.

Consider the subset

of $Z(P- \varepsilon )$ , i.e., the set of points in $Z(P - \varepsilon )$ at which the gradient of P is collinear to $x_1$ -axis. At any point $\mathbf {x} \in G \cap \mathcal {C}_i$ the gradient of P projects injectively on the line, normal to $\mathrm {proj}(Z(P - \varepsilon ) \cap \mathcal {C}_i)$ at $\mathrm {proj} (\mathbf {x})$ . It may happen, that $\mathrm {proj}\ (G \cap \mathcal {C}_i)$ has infinite number of points (in case when this set consists of critical points of the Gauss map of the manifold $\mathrm {proj}\ (Z(P - \varepsilon ) \cap \mathcal {C}_i)$ ). Sard’s theorem implies that using a generic rotation of coordinates $x_1, \ldots , x_k$ we can guarantee that $\mathrm {proj}\ G \cap \mathcal {C}_i$ contains only regular points of the Gauss map, and therefore, is finite. As a result, the preimage of this set under the embedding, $G \cap \mathcal {C}_i$ , is also finite.

The set $G \cap \mathcal {C}_i$ is defined by the system of equations

$$\begin{align*}\left\{\text{graph of }\mathbf{f}_i, P - \varepsilon = \frac{\partial P}{\partial x_2} = \ldots = \frac{\partial P}{\partial x_{k}} = 0\right\} .\end{align*}$$

Recall from Lemma 3.1 that the graph of $\mathbf {f}_i$ , which is just $\mathcal {C}_i$ , is an intersection of an open set, say $U_i$ , and a set of solutions of a system of Pfaffian equations. The domain of these Pfaffian equations must obviously be larger than $U_i$ . Thus we use Corollary 2.15 to obtain an upper bound on $|G \cap \mathcal {C}_i|$ ; by doing so, we are able to conclude that the number of solutions of this system is at most $C"D^{k+r}$ , for some constant $C" = C"(n, k, m, \alpha , \beta , r)$ that comes from applying Corollary 2.15.

As argued above, $Z(P - \varepsilon ) \cap \alpha _j$ , for all $j \in [A]$ , is compact; in other words, each connected compoment in $\{\alpha _j\}_{j \in A}$ properly contains at least one component of $Z(P - \varepsilon )$ . Thus at least one point of $G \cap \mathcal {C}_i$ is sure to be present in each of $\{\alpha _i\}_{i \in A}$ . This means that

$$\begin{align*}|A| \le C"D^{k+r}.\end{align*}$$

The quantitative results above would be exactly the same if we looked at $P + \varepsilon $ , for some $\varepsilon < 0$ , instead of $P - \varepsilon $ , so $C"D^{k+r}$ is an upper bound on $\left |[t] \setminus A\right |$ as well. This means that $2C"D^{k+r}$ is an upper bound on t, and thus by (3.2), the proof is complete by choosing $C'$ appropriately.

Using Theorem 1.6, we are able to prove Theorem 1.4. In fact, we are able to prove Theorem 3.2, which is a generalization of Theorem 1.4.

Theorem 3.2 (Polynomial partitioning of collections of semi-Pfaffian sets - generalization of Theorem 1.4-(I))

Let $m \ge 1$ be an integer, and for each $i \in [m]$ , let $\Gamma _i$ be a finite set of $k_i$ -dimensional semi-Pfaffian sets all in $\mathbb {R}^n$ , where each $\mathcal {X} \in \Gamma _i$ is defined by at most $h_i$ Pfaffian functions each of degree at most $\beta _i$ , chain-degree at most $\alpha _i$ , and order at most $r_i$ . Then for any $D \ge 1$ , there exists a nonzero polynomial $P \in \mathbb {R}[X_1, \ldots , X_n]$ of degree at most D, and constants $C_i = C_i(n, k_i, h_i, \alpha _i, \beta _i, r_i)$ such that

  1. (I) for all $i \in [m]$ with $k_i> 0$ , each connected component of $\mathbb {R}^n \setminus Z(P)$ intersects at most $C_im\frac {|\Gamma _i|}{D^{n-k_i-r_i}}$ elements of $\Gamma _i$ , and

  2. (II) for all $i \in [m]$ with $k_i = 0$ , each connected component of $\mathbb {R}^n \setminus Z(P)$ intersects at most $C_im\frac {|\Gamma _i|}{D^n}$ elements of $\Gamma _i$ .

Theorem 3.2 can be proved by following the exact same strategy as in [Reference Blagojević, Blagojević and Ziegler13], but by using Theorem 1.6 at the appropriate place. We reproduce the main part of the argument in [Reference Blagojević, Blagojević and Ziegler13] in Appendix A to make our document self-contained.

Proof of Theorem 1.4

Part (I) is just the special case of Theorem 3.2 with $m=1$ .

For Part (II), recall that we assume that all the Pfaffian functions involved are defined with respect to the same Pfaffian chain $(q_1, \ldots , q_r)$ . To every $\mathcal {X}\in \Gamma $ we associate a semi-Pfaffian set $\mathcal {X}'\subset \mathbb {R}^{n+r}$ as follows. Let $\mathcal {X}$ be defined by Pfaffian functions $P_i(\mathrm {x}) = Q_i(\mathrm {x}, q_1(\mathrm {x}), \ldots , q_r(\mathrm {x}))$ , $i \in [m]$ . To $\mathcal {X}$ , we associate $\mathcal {X}'\subset \mathbb {R}^{n+r}$ , where $\mathcal {X}'$ is the set defined by the polynomials $Q_i\in \mathbb {R}[X_1, \ldots , X_n, Y_1, \ldots , Y_r]$ intersected with $Z(y_1-q_1(\mathrm {x}), \ldots , y_r-q_r(\mathrm {x}))$ (that is, the intersection of a semi-algebraic set with the graph of the Pfaffian chain). In other words, letting $\mathcal {G} \subseteq \mathbb {R}^{n+r}$ denote the graph of the map

$$\begin{align*}\mathbb{R}^n \ni \mathrm{x} \mapsto (q_1(\mathrm{x}), \ldots, q_r(\mathrm{x})) \in \mathbb{R}^{r}, \end{align*}$$

it is clear that $\mathcal {X}'$ is obtained by restricting the domain of $\mathcal {G}$ to $\mathcal {X}$ .

Let $\Gamma '$ denote the collection of the associated semi-algebraic sets $\mathcal {X}'$ . Each $\mathcal {X}'$ has dimension k, so we can apply Part (I) to infer the existence of a polynomial $P\in \mathbb {R}[X_1, \ldots , X_n, Y_1, \ldots , Y_r]$ of degree at most D such that each connected component of $\mathbb {R}^{n+r}\setminus Z(P)$ intersects at most $C'D^{k+r-(n-r)}|\Gamma '|=C'D^{k-n}|\Gamma |$ elements of $\Gamma '$ . Specifically, since all $\Gamma ' \ni \mathcal {X}' \subseteq \mathcal {G}$ , the guarantee of Part 1.4 is equivalent to saying that each connected component of $\mathcal {G} \setminus (\mathcal {G} \cap Z(P))$ contains at most $C'D^{k-n}|\Gamma |$ elements of $\Gamma '$ .

Define $P'(\mathrm {x}) := P(\mathrm {x}, q_1(\mathrm {x}), \ldots , q_r(\mathrm {x})) \in \mathrm {Pfaff}_n(\alpha , \beta , r)$ . $\mathcal {G}$ is homotopic to an open set in $\mathbb {R}^n$ , thus the projection of $\mathcal {G}$ onto $\mathbb {R}^n$ carries each $\mathcal {X}'$ to $\mathcal {X}$ , and also carries $\mathcal {G} \cap Z(P)$ to $Z(P')$ . In other words, intersections of the various $\mathcal {X} \in \Gamma $ with connected components of $\mathbb {R}^n\setminus Z(P')$ bijectively correspond to intersections of the corresponding $\mathcal {X}' \in \Gamma '$ with connected components of $\mathcal /{G} \setminus (\mathcal {G} \cap Z(P))$ . Since each connected component of $\mathcal {G} \setminus (\mathcal {G} \cap Z(P))$ contains at most $C'D^{k-n}|\Gamma |$ elements of $\Gamma '$ , the bijectivity of the projection from $\mathcal {G}$ onto $\mathbb {R}^n$ ensures that each connected component of $\mathbb {R}^n \setminus Z(P')$ contains at most $C'D^{k-n}|\Gamma |$ elements of $\Gamma $ .

It remains to prove that $P'$ is nonzero. By the guarantee of Theorem 1.4-(I), P is nonzero. Since $P \in \mathbb {R}[X_1, \ldots , X_n, Y_1, \ldots , Y_r]$ , we can also consider that $P \in \mathbb {R}[Y_1, \ldots , Y_r][X_1, \ldots , X_n]$ , i.e. that P consists of monomials in the variables $X_1, \ldots X_n$ with coefficients from $\mathbb {R}[Y_1, \ldots , Y_r]$ . Since P is nonzero, at least one of the coefficients is nonzero, and since we have assumed that $\vec {q}$ is an algebraically independent chain, we have that no coefficients of P will become identically zero when we substitute all $Y_i$ by $q_i$ . Thus we have that $P'$ will have at least one nonzero coefficient, proving that $P'$ is nonzero.

By following the exact same strategy as in the proof of Theorem 1.4 - (II) and by using Theorem 3.2 instead of Theorem 1.4 - (I), we get the following theorem, whose proof we only sketch.

Theorem 3.3 (Pfaffian partitioning of collections of semi-Pfaffian sets, - generalization of Theorem 1.4-1.4)

Let $\vec {q} = (q_1, \ldots , q_r)$ be an algebraically independent Pfaffian chain of order r and chain-degree $\alpha $ . Let $m \ge 1$ be an integer, and for each $i \in [m]$ , let $\Gamma _i$ be a finite set of $k_i$ -dimensional semi-Pfaffian sets all in $\mathbb {R}^n$ , where each $\mathcal {X} \in \Gamma _i$ is defined by at most $h_i$ functions from $\mathrm {Pfaff}_n(\beta _i, \vec {q})$ . For any $D \ge 1$ , there is a nonzero Pfaffian function $P' \in \mathrm {Pfaff}_n(D, \vec {q})$ , and constants $C_i = C_i(n, k_i, h_i, \alpha , \beta _i, r)$ such that

  1. (I) for all $i \in [m]$ with $k_i> 0$ , each connected component of $\mathbb {R}^n \setminus Z(P')$ intersects at most $C_im\frac {|\Gamma _i|}{D^{n - k_i}}$ elements of $\Gamma _i$ , and

  2. (II) for all $i \in [m]$ with $k_i = 0$ , each connected component of $\mathbb {R}^n \setminus Z(P')$ intersects at most $C_im\frac {|\Gamma _i|}{D^{n+r}}$ elements of $\Gamma _i$ .

Proof sketch of Theorem 3.3

Just as in the proof of Theorem 1.4-(II), we lift from $\mathbb {R}^n$ each collection of semi-Pfaffian sets $\Gamma _i$ to $\mathbb {R}^{n+r}$ to form $\{\Gamma _i'\}_{i \in [m]}$ where each $\Gamma _i'$ is a collection of $k_i$ -dimensional semi-Pfaffian sets in $\mathbb {R}^{n+r}$ . We now apply Theorem 3.2 to find a polynomial $P \in \mathbb {R}[X_1, \ldots , X_n, Y_1, \ldots , Y_r]$ of degree at most D such that

  1. (I) for all $i \in [m]$ with $k_i> 0$ , each connected component of $\mathbb {R}^{n+r} \setminus Z(P)$ intersects at most $\sim \frac {|\Gamma _i'|}{D^{n-k_i}}$ elements of $\Gamma _i'$ , and

  2. (II) for all $i \in [m]$ with $k_i = 0$ , each connected component of $\mathbb {R}^{n+r} \setminus Z(P)$ intersects at most $\sim \frac {|\Gamma _i'|}{D^{n+r}}$ elements of $\Gamma _i'$ .

Next, again as in the proof of Theorem 1.4-(II), we turn P into a Pfaffian function

$$\begin{align*}P'(\mathrm{x}) = P(\mathrm{x}, q_1(\mathrm{x}), \ldots, q_r(\mathrm{x})) \in \mathrm{Pfaff}_n(D, \vec{q}), \end{align*}$$

and $P'$ satisfies the required guarantees on the collection $\{\Gamma _i\}_{i \in [m]}$ .

4 Applications

In this section, we derive some applications of our main theorems. Sections 4.1 and 4.2 contain definitions and theorem statements, Section 4.3 contains some discussion on finer technical details with regard to the results, and Section 4.4 contains proofs of the stated theorems.

4.1 Szemerédi-Trotter type theorems

Given a finite set $\mathcal {P}$ of points in $\mathbb {R}^n$ , and a finite set $\Gamma $ of subsets of $\mathbb {R}^n$ , the set of incidences between $\mathcal {P}$ and $\Gamma $ is defined as

$$ \begin{align*} I(\mathcal{P}, \Gamma) := \left\{(p, \gamma) \in \mathcal{P} \times \Gamma \boldsymbol{:} p \in \gamma\right\}. \end{align*} $$

Also, the incidence graph of $\Gamma $ and $\mathcal {P}$ , denoted $G_{\mathcal {P}, \Gamma }$ , is a bipartite graph $G_{\mathcal {P}, \Gamma } = (V_{\mathcal {P}} \cup V_{\Gamma }, E)$ , where the vertices $V_{\mathcal {P}}$ corresponding to points in $\mathcal {P}$ and the vertices $V_{\Gamma }$ correspond to elements of $\Gamma $ , and $(v_p, v_{\gamma }) \in E$ if $p \in \gamma $ .

The classical Szemerédi-Trotter Theorem [Reference Szemerédi and Trotter62] proves a bound of $O(|\mathcal {P}|^{2/3}|\Gamma |^{2/3}+|\mathcal {P}|+|\Gamma |)$ on the number of incidences $|I(\mathcal {P}, \Gamma )|$ when $\Gamma \subseteq \mathbb {R}^2$ consists purely of lines. Under certain conditions on the incidence graph (see [Reference Pach and Sharir51, Reference Clarkson, Edelsbrunner, Guibas, Sharir and Welzl19] or [Reference Sheffer58, Chapter 3]), this bound has been generalized to the case of algebraic curves. In [Reference Solymosi and Tao60], incidences between points and higher dimensional varieties are studied under specific nondegeneracy assumptions; we refer to [Reference Sheffer58] for many more such theorems when $\Gamma $ contains algebraic sets. Also, as mentioned earlier, [Reference Basu and Raz9, Reference Chernikov, Galvin and Starchenko15] derive o-minimal/distal Szemerédi-Trotter type theorems.

Let $K_{s,t}$ denote the complete bipartite graph with s vertices on one side and t vertices on the other. The following Theorem counts incidences between points and Pfaffian curves in $\mathbb {R}^n$ under some nondegeneracy assumptions.

Theorem 4.1. Let $m, \alpha , r, \beta $ , $n \ge 3$ , and $s, t \ge 2$ , be integers, and let $\varepsilon> 0$ . Then there exist constants $c_j = c_j(\varepsilon , n, m, \alpha , \beta , r, s, t)$ , for $1 < j < n$ , such that the following holds. Let $\vec {q}$ be an algebraically independent Pfaffian chain of order r and chain-degree $\alpha $ . Let $\mathcal {P} \subseteq \mathbb {R}^n$ be a set of points, and let $\Gamma \subseteq \mathbb {R}^n$ be a set of distinct irreducible Pfaffian curves such that each $\gamma \in \Gamma $ is defined by at most m Pfaffian functions from $\mathrm {Pfaff}_n(\beta , \vec {q})$ . Suppose that $G_{\mathcal {P}, \Gamma }$ does not contain any copy of $K_{s,t}$ . Also assume that, for all $1 < j < n$ , every j-dimensional irreducible Pfaffian set defined by Pfaffian functions from $\mathrm {Pfaff}_n(c_j, \vec {q})$ contains at most $\theta _j$ curves of $\Gamma $ . Then there exist constants $C_1 = C_1(\varepsilon , n, m, \alpha , \beta , r, s, t)$ and $C_2 = C_2(\varepsilon , n, m, \alpha , \beta , r, s, t)$ such that

$$\begin{align*}|I(\mathcal{P}, \Gamma)| \le C_1 |\mathcal{P}|^{\frac{s(r+1)}{s(n+r)-n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r) -n + 1}} + |\mathcal{P}|\sum_{j=2}^{n-1} \theta_j + C_2 (|\Gamma| + |\mathcal{P}|). \end{align*}$$

Corollary 4.2 (of Theorem 4.1)

Let $\vec {q}$ be an algebraically independent Pfaffian chain of order r and chain-degree $\alpha $ . Let $\mathcal {P} \subseteq \mathbb {R}^2$ be a set of points, and let $\Gamma \subseteq \mathbb {R}^2$ be a set of distinct irreducible Pfaffian curves such that each $\gamma \in \Gamma $ is defined by at most m Pfaffian functions from $\mathrm {Pfaff}_n(\beta , \vec {q})$ . Suppose that $G_{\mathcal {P}, \Gamma }$ does not contain any copy of $K_{s,t}$ . Then, for any $\varepsilon> 0$ , there exist constants $C_1 = C_1(\varepsilon , m, \alpha , \beta , r, s, t)$ and $C_2 = C_2(\varepsilon , m, \alpha , \beta , r, s, t)$ such that

$$\begin{align*}|I(\mathcal{P}, \Gamma)| \le C_1 \left(|\mathcal{P}|^{\frac{s(r+1)}{s(r+2) - 1} + \varepsilon}|\Gamma|^{\frac{(s-1)(r+2)}{s(r+2) - 1}}\right) + C_2 \left(|\mathcal{P}| + |\Gamma|\right). \end{align*}$$

Proof of Corollary 4.2

Immediate by setting $n=2$ and $\theta _j = 0$ in Theorem 4.1

Remark 4.3. Theorem 4.1 works under the assumption that low degree Pfaffian sets do not contain too many curves from $\Gamma $ . This assumption matches the assumption of the main incidence theorems in [Reference Sharir, Sheffer and Solomon57] (henceforth this will be called the SSS assumption), where they study incidences between points and algebraic curves in $\mathbb {R}^{n}$ when $n \ge 3$ .

In the algebraic case, when the ambient space is $\mathbb {R}^2$ , an SSS type assumption can be avoided simply because it is well known that an algebraic curve in $\mathbb {R}^2$ of degree D contains at most D irreducible components. However, in the Pfaffian case, the question of counting the number irreducible components becomes complicated. Counting irreducible components has been achieved only in restricted cases; for example, [Reference Roy and Vorobjov54] study irreducible (over the real algebraic numbers) components of exponential polynomials. However, using Corollary 2.18 where we obtain a coarse bound on the number of irreducible components, we are able to avoid the SSS assumption in the statement of Corollary 4.2.

4.2 Pfaffian Joints

Definition 4.4 (Joints)

Let $m, n \ge 2$ . For $i \in [m]$ , suppose $\Gamma _i$ is a set of $k_i$ -dimensional semi-Pfaffian sets in $\mathbb {R}^n$ such that $\sum _{i=1}^m k_i = n$ . A point $x \in \mathbb {R}^n$ is a joint if

  1. (I) for each $i \in [m]$ , there exists $\gamma _i \in \Gamma _i$ such that x is a smooth point of $\gamma _i$ , and

  2. (II) the tangent spaces of $\gamma _i$ at x for all i span $\mathbb {R}^n$ .

The following theorem establishes an upper bound on the number of joints formed by n sets of Pfaffian curves in $\mathbb {R}^n$ .

Theorem 4.5. Let $\vec {q}$ be an algebraically independent Pfaffian chain of order r and chain-degree $\alpha $ . Suppose there are $n \ge 3$ sets of distinct irreducible Pfaffian curves $\{\Gamma _i\}_{i \in [n]}$ in $\mathbb {R}^n$ such that for all $i \in [n]$ , each $\mathcal {X}_i \in \Gamma _i$ is defined by at most m Pfaffian functions from $\mathrm {Pfaff}_n(\beta , \vec {q})$ . Then for any $\varepsilon> 0$ , there exists a constant $C = C(\varepsilon , n, m, \alpha , \beta , r)$ such that the number of joints is bounded by

$$\begin{align*}C\cdot \min_{j \in [n]} \left\{ |\Gamma_j|^{\varepsilon}\right\} \cdot \prod_{i = 1}^n |\Gamma_i|^{\max\left\{\frac{n + r}{n(n - 1)}, \frac{2}{n+1}\right\}}. \end{align*}$$

Remark 4.6. Using Lemma 3.1, one can remove the requirement of irreducibility of the Pfaffian curves in the hypothesis of Theorem 4.1, and Corollary 4.2, and make it work for arbitrary semi-Pfaffian curves. The same holds for Theorem 4.5 as well.

4.3 Discussion

4.3.1 Nondegeneracy assumptions

In our Szemerédi-Trotter type Theorems, we assume that the incidence graph does not contain any copy of $K_{s,t}$ . Suppose $\Gamma $ in Theorem 4.1 was just a set of lines, then it is true that $K_{2,2} \not \subseteq G_{\mathcal {P}, \Gamma }$ . Thus Theorem 4.1 generalizes the classical version of the Szemerédi-Trotter theorem [Reference Szemerédi and Trotter62]. Similarly, if $\Gamma $ is purely a set of unit circles, then $G_{\mathcal {P}, \Gamma }$ contains no copy of $K_{2,3}$ , or if $\Gamma $ is a set of arbitrary circles, $G_{\mathcal {P}, \Gamma }$ contains no copy of $K_{3,2}$ .

In general, if $\Gamma $ is a general set of varieties, it is quite natural to require some kind of nondegeneracy assumptions on $\mathcal {P}$ and $\Gamma $ to prove nontrivial incidence bounds. For example, in [Reference Solymosi and Tao60], a bound on incidences between points and higher dimensional varieties is proved assuming some axioms that they call pseudoline axioms – i.e. axioms which ensure that $\Gamma $ behaves somewhat like a set of lines. [Reference Pach and Sharir51, Theorem 1.1] proves an incidence bound for points and simple curves in $\mathbb {R}^2$ with an assumption equivalent to not having $K_{s,t}$ inside $G_{\mathcal {P}, \Gamma }$ , although it is worded slightly differently. [Reference Sharir, Sheffer and Solomon57, Theorem 1.4 and Theorem 1.5] prove incidence bounds for points and curves in $\mathbb {R}^d$ , $d \ge 3$ , by making an assumption that ensures that no partitioning hypersurface can contain too many elements of $\Gamma $ ; we include the same assumption in Theorem 4.1. [Reference Basu and Raz9] prove an o-minimal Szemerédi-Trotter theorem for points and plane curves assuming there is no copy of $K_{2, t}$ for some fixed t.

An example of a degenerate case is where $\Gamma $ is a set of planes all intersecting in the same line (one might visualise a partially open book where the pages are the planes, and the spine of the book is the common line of intersection), and all $\mathcal {P}$ lie on the line. Then the number of incidences is exactly $|P||\Gamma |$ , which is the trivial upper bound.

4.3.2 Degree of partitioning polynomial or partitioning Pfaffian function

When applying polynomial partitioning, or in our case, perhaps Pfaffian partitioning, one needs to decide what the degree of the polynomial or Pfaffian function must be. In incidence geometry literature, there is a dichotomy in this choice - high degree partitioning vs low/constant degree partitioning.

For instance, the sharp version of the original Szemerédi-Trotter theorem, where we count incidences between points and lines in $\mathbb {R}^2$ , can be proved using polynomial partitioning where the degree of the partitioning polynomial P is high, i.e. the degree depends on the cardinalities of $\mathcal {P}$ and $\Gamma $ [Reference Kaplan, Matoušek and Sharir36]. Since the degree of P is high, the number of points/lines in each connected component of $\mathbb {R}^2 \setminus Z(P)$ is small; this allows us to use the naive bound of Theorem 4.7 to count the number of incidences in each connected component, and subsequently amplify it. However, the higher degree of P in turn means that the number of incidences on $Z(P)$ becomes higher. While there is no trouble in $\mathbb {R}^2$ , using the same strategy in higher dimensions causes trouble.

[Reference Solymosi and Tao60] introduced the idea of partitioning with a constant degree polynomial. But for a loss in power of $\varepsilon $ , this strategy allowed them to prove sharp bounds on incidences in higher dimensions between points and high-dimensional varieties. In this case, while counting incidences on the zero set of the partitioning polynomial becomes easier, one can no longer use the naive bound of Theorem 4.7 to count incidences in connected components; one has to proceed by induction with a strong enough hypothesis. In fact, we lose a factor $\varepsilon $ in the exponent precisely to close the induction.

In Theorem 4.1 and Theorem 4.5, we use constant degree partitioning, and particularly, we use constant degree Pfaffian partitioning. While counting incidences on Pfaffian sets is indeed harder than on algebraic sets, the fact that the degree of the partitioning Pfaffian function is constant assuages this problem.

4.3.3 Projections of Pfaffian sets

While one might hope to generalize Theorem 4.1 to incidences between points and higher dimensional semi-Pfaffian sets, and generalize Theorem 4.5 to counting joints between higher dimensional Pfaffian sets, there is one common issue that stymies both directions, i.e. the issue of projections. While semi-algebraic sets project into semi-algebraic sets by the Tarski-Seidenberg theorem, projections of semi-Pfaffian sets, called sub-Pfaffian sets, need not be semi-Pfaffian. Such an example is provided in [Reference Osgood50].

For an example of how this issue manifests, suppose we wished to count incidences between points and two-dimensional semi-Pfaffian sets in $\mathbb {R}^3$ . In this case, counting incidences on the zero set of the partitioning Pfaffian function becomes difficult; we cannot induct using Corollary 4.2 because two-dimensional semi-Pfaffian sets which do not lie on the partitioning Pfaffian set intersect in one-dimensional Pfaffian sets, and we cannot project these one-dimensional intersections down to $\mathbb {R}^2$ because of the problem mentioned above.

One way around this problem could be to develop multilevel polynomial/Pfaffian partitioning theorems, which is a challenging problem with many works devoted to it [Reference Matoušek and Patáková46, Reference Basu and Sombra10, Reference Fox, Pach, Sheffer, Suk and Zahl24, Reference Walsh66]. However, this is out of the scope of our work, and we leave it for future work.

4.4 Proofs of Theorem 4.1 and Theorem 4.5

4.4.1 Proof of Szemerédi-Trotter type theorem

To prove our Szemerédi-Trotter type theorems, we make use of the following naive bound on incidences between points and an arbitrary set of sets in $\mathbb {R}^n$ . In fact, this works not just over $\mathbb {R}$ , but any field.

Theorem 4.7 (Kővári-Sós-Turán Theorem [Reference Matoušek45, Theorem 4.5.2])

Let $\mathcal {P}$ be a set of points in $\mathbb {R}^n$ and let $\Gamma $ be a set of subsets of $\mathbb {R}^n$ . Suppose that $G_{\mathcal {P}, \Gamma }$ does not contain any copy of $K_{s,t}$ . Then there exists a constant $C = C(s, t)$ such that

$$\begin{align*}|I(\mathcal{P}, \Gamma)| \le C \cdot \min\left(|\mathcal{P}||\Gamma|^{\frac{s-1}{s}} + |\Gamma|, |\mathcal{P}|^{\frac{t-1}{t}}|\Gamma| + |\mathcal{P}|\right).\end{align*}$$

Proof of Theorem 4.1

In general, various intermediary constants that depend on $\varepsilon $ , n, m, $\alpha $ , $\beta $ , r, s, and t, shall all be denoted by $\kappa $ without disambiguation; for example, $\kappa ^2$ in some step will be replaced with $\kappa $ in the next step. Only when we need to disambiguate, we will use $\kappa _i$ . Later in the proof, we will use Theorem 3.3 to partition $\mathcal {P}$ and $\Gamma $ simultaneously; to distinguish, we let $C_0$ denote the maximum of all constants in Theorem 3.3. Also, let $C'$ denote the constant of Theorem 4.7.

We will proceed by induction on $|\mathcal {P}| + |\Gamma |$ . The base case of $|\mathcal {P}| + |\Gamma | = \kappa $ can be handled by using Theorem 4.7, i.e.,

$$\begin{align*}|I(\mathcal{P}, \Gamma)| \le C'(\kappa^{\frac{s-1}{s}}\kappa + \kappa) \le \kappa_1, \end{align*}$$

thus setting $C_2 \ge \kappa _1$ proves the base case.

For the induction step, we use Theorem 3.3 to infer the existence of $P \in \mathrm {Pfaff}_n(D, \vec {q})$ , with $D = D(\varepsilon , n, m, \alpha , \beta , r, s, t)$ to be specified later, such that $\mathbb {R}^n \setminus Z(P)$ has connected components $\mathcal {C}_1, \ldots , \mathcal {C}_{\alpha }$ satisfying, for all $i \in [\alpha ]$ ,

$$\begin{align*}|\mathcal{P} \cap \mathcal{C}_i| \le C_0\frac{|\mathcal{P}|}{D^{n+r}}, \qquad \text{and} \qquad \left|\{\gamma \in \Gamma \boldsymbol{:} \gamma \cap \mathcal{C}_i \neq \emptyset\}\right| \le C_0\frac{|\Gamma|}{D^{n-1}}. \end{align*}$$

Setting $D \ge C_0^{\frac {1}{n+r}} + 1$ , we have that $C_0\frac {|\mathcal {P}|}{D^{n+r}} < |\mathcal {P}|$ ; thus we can use the induction hypothesis to obtain that the number of incidences in $\mathcal {C}_i$ is bounded as

(4.1) $$ \begin{align} &|I(\mathcal{P}, \Gamma) \cap \mathcal{C}_i| \nonumber \\ &\le C_1 \left(\left(\frac{C_0|\mathcal{P}|}{D^{n+r}}\right)^{\frac{s(r+1)}{s(n+r)-n+1} + \varepsilon}\left(\frac{C_0|\Gamma|}{D^{n-1}}\right)^{\frac{(s-1)(n+r)}{s(n+r) -n + 1}} \right) + |\mathcal{P} \cap \mathcal{C}_i|\sum_{j=2}^{n-1} \theta_j \nonumber \\ & \qquad \qquad \qquad + C_2\left(|\mathcal{P} \cap \mathcal{C}_i| + \frac{C_0|\Gamma|}{D^{n-1}}\right) \nonumber \\ &= C_1 C_0^{\frac{s(r+1) + (s-1)(n+r)}{s(n+r) - n + 1} + \varepsilon} \left(\frac{|\mathcal{P}|}{D^{n+r}}\right)^{\frac{s(r+1)}{s(n+r) - n + 1} + \varepsilon}\left(\frac{|\Gamma|}{D^{n-1}}\right)^{\frac{(s-1)(n+r)}{s(n+r)-n + 1}} + |\mathcal{P} \cap \mathcal{C}_i|\sum_{j=2}^{n-1} \theta_j \nonumber \\ & \qquad \qquad \qquad+ C_2\left(|\mathcal{P} \cap \mathcal{C}_i| + \frac{C_0|\Gamma|}{D^{n-1}}\right) \nonumber \\ &= \kappa C_1 \left(\frac{|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n + 1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n + 1}}}{D^{\frac{s(n+r)(r+1) + (s-1)(n+r)(n-1)}{s(n+r) - n + 1} + (n+r)\varepsilon}}\right) + |\mathcal{P} \cap \mathcal{C}_i|\sum_{j=2}^{n-1} \theta_j \nonumber \\ &\qquad \qquad \qquad + C_2\left(|\mathcal{P} \cap \mathcal{C}_i| + \frac{C_0|\Gamma|}{D^{n-1}}\right) \nonumber \\ &= \kappa C_1 D^{-(n+r)(1 + \varepsilon)}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n+1}} + |\mathcal{P} \cap \mathcal{C}_i|\sum_{j=2}^{n-1} \theta_j \nonumber \\ &\qquad \qquad \qquad + C_2\left(|\mathcal{P} \cap \mathcal{C}_i| + \frac{C_0|\Gamma|}{D^{n-1}}\right). \end{align} $$

If $|P| < |\Gamma |^{\frac {1}{s}}$ , then from Theorem 4.7 we’d have that

$$\begin{align*}|I(\mathcal{P}, \Gamma)| \le C'(|\mathcal{P}||\Gamma|^{\frac{s-1}{s}} + |\Gamma|) < 2C'|\Gamma|, \end{align*}$$

and as long as $C_2 \ge 2C'$ , we would be done. Thus without loss of generality, we can assume

(4.2) $$ \begin{align} |\Gamma|^{\frac{1}{s}} \le |\mathcal{P}|. \end{align} $$

Now, assuming (4.2), we have

(4.3) $$ \begin{align} |\Gamma| = |\Gamma|^{\frac{(s-1)(n+r)}{s(n+r) - n + 1}}|\Gamma|^{\frac{(r+1)}{s(n+r) - n+1}} \le |\Gamma|^{\frac{(s-1)(n+r)}{s(n+r) - n+1}}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n + 1}}. \end{align} $$

Noting that by Theorem 1.6 (and also [Reference Gabrielov and Vorobjov25, Theorem 3.4]), we have that $\mathbb {R}^n \setminus Z(P)$ has $\alpha \le \kappa D^{n+r}$ connected components, we estimate the total number of incidences in all the connected components of $\mathbb {R}^n \setminus Z(P)$ as follows:

(4.4) $$ \begin{align} &\left|I(\mathcal{P}, \Gamma) \cap \left(\bigcup_{i=1}^{\alpha} \mathcal{C}_i\right)\right| \nonumber \\ &= \sum_{i=1}^{\alpha} \left|I(\mathcal{P}, \Gamma) \cap \mathcal{C}_i\right| \nonumber \\ &\le \sum_{i=1}^{\alpha} \left(\kappa C_1 D^{-(n+r)(1 + \varepsilon)}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n+1}}\right) \nonumber \\ &\qquad \qquad \qquad + \sum_{i=1}^{\alpha} \left(|\mathcal{P} \cap \mathcal{C}_i|\sum_{j=2}^{n-1} \theta_j + C_2\left(|\mathcal{P} \cap \mathcal{C}_i| + \frac{C_0|\Gamma|}{D^{n-1}}\right)\right) \qquad\qquad\qquad\qquad\qquad(\text{by}\ (4.1)) \nonumber \\ &= \alpha \cdot \kappa C_1 D^{-(n+r)(1 + \varepsilon)}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n+1}} \nonumber \\ &\qquad \qquad \qquad + \sum_{i=1}^{\alpha} \left(|\mathcal{P} \cap \mathcal{C}_i|\sum_{j=2}^{n-1} \theta_j + C_2\left(|\mathcal{P} \cap \mathcal{C}_i| + \frac{C_0|\Gamma|}{D^{n-1}}\right)\right) \nonumber \\ &\le \kappa C_1 D^{-(n+r)\varepsilon}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n+1}} + \left|\mathcal{P} \cap \left(\bigcup _{i=1}^{\alpha} \mathcal{C}_i\right)\right|\sum_{j=2}^{n-1} \theta_j \nonumber \\ &\qquad \qquad \qquad + C_2\left|\mathcal{P} \cap \left(\bigcup_{i=1}^{\alpha} \mathcal{C}_i \right)\right| + \alpha C_2\left(\frac{C_0|\Gamma|}{D^{n-1}}\right) \nonumber \\ &\le \kappa_2 C_1 D^{-(n+r)\varepsilon}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n+1}} + \left|\mathcal{P} \cap \left(\bigcup _{i=1}^{\alpha} \mathcal{C}_i\right)\right|\sum_{j=2}^{n-1} \theta_j \nonumber \\ &\qquad \qquad \qquad + C_2\left|\mathcal{P} \cap \left(\bigcup_{i=1}^{\alpha} \mathcal{C}_i \right)\right| + \kappa D^{r+1}C_2|\Gamma| \nonumber \\ &\le \kappa_2 C_1 D^{-(n+r)\varepsilon}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n+1}} + \left|\mathcal{P} \cap \left(\bigcup _{i=1}^{\alpha} \mathcal{C}_i\right)\right|\sum_{j=2}^{n-1} \theta_j \nonumber \\ &\qquad \qquad \qquad + C_2\left|\mathcal{P} \cap \left(\bigcup_{i=1}^{\alpha} \mathcal{C}_i \right)\right| + \kappa_3 D^{r+1}C_2|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r) - n+1}}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n + 1}} \qquad(\text{by}\ (4.3)) \nonumber \\ &\le \frac{C_1}{4}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n+1}} + \left|\mathcal{P} \cap \left(\bigcup _{i=1}^{\alpha} \mathcal{C}_i\right)\right|\sum_{j=2}^{n-1} \theta_j + C_2 \left|\mathcal{P} \cap \left(\bigcup _{i=1}^{\alpha} \mathcal{C}_i\right)\right| , \end{align} $$

where the last line holds as long $D \ge (8\kappa _2)^{\frac {1}{(n+r)\varepsilon }}$ and $C_1 \ge 8\kappa _3D^{r+1}C_2$ .

Define $\Gamma ' := \{\gamma \in \Gamma \boldsymbol {:} \gamma \subseteq Z(P)\}$ . Apply Corollary 2.17 to decompose $Z(P)$ , and let $Z_1$ denote the union of all the irreducible $1$ -dimensional components in the decomposition. Define $\Gamma _1' := \{\gamma \in \Gamma \boldsymbol {:} \gamma \subseteq Z_1\} \subseteq \Gamma '$ . Now, by Corollary 2.18, we have that the number of irreducible 1-dimensional components of $Z_1$ is at most $\kappa D^{\kappa }$ , thus we have that

(4.5) $$ \begin{align} |I(\mathcal{P} \cap Z(P), \Gamma_1')| \le |\mathcal{P} \cap Z(P)| \cdot |\Gamma_1'| \le \kappa |\mathcal{P} \cap Z(P)| D^{\kappa} = \kappa_4|\mathcal{P} \cap Z(P)|. \end{align} $$

For all $1 < j < n$ , let $Z_j$ denote the union of all the irreducible j-dimensional components of $Z(P)$ . Correspondingly, define $\Gamma _j' := \{\gamma \in \Gamma \boldsymbol {:} \gamma \subseteq Z_j\} \subseteq \Gamma '$ . By assumption, we have

(4.6) $$ \begin{align} |I(\mathcal{P} \cap Z(P), \Gamma_j')| \le |\mathcal{P} \cap Z(P)| \cdot |\Gamma_j'| \le \theta_j |\mathcal{P} \cap Z(P)|. \end{align} $$

Also, by Lemma 3.1 and Corollary 2.15, we have that each $\gamma \in \Gamma \setminus \Gamma '$ intersects $Z(P)$ in at most $\kappa D^{r+1}$ points, thus

(4.7) $$ \begin{align} |I(\mathcal{P} \cap Z(P), \Gamma \setminus \Gamma')| \le \kappa|\Gamma \setminus \Gamma'|D^{r+1} \le \kappa_5 |\Gamma|. \end{align} $$

Finally, from (4.4), (4.5), (4.7),and (4.7), we deduce

$$ \begin{align*} &|I(\mathcal{P}, \Gamma)| \nonumber \\ &= \left|I(\mathcal{P}, \Gamma) \cap \left(\bigcup _{i=1}^{\alpha} \mathcal{C}_i\right)\right| + \sum_{j=1}^{n-1} |I(\mathcal{P} \cap Z(P), \Gamma_j')| + |I(\mathcal{P} \cap Z(P), \Gamma \setminus \Gamma')| \nonumber \\ &\le \frac{C_1}{4}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n+1}} + \left|\mathcal{P} \cap \left(\bigcup _{i=1}^{\alpha} \mathcal{C}_i\right)\right|\sum_{j=2}^{n-1} \theta_j + C_2 \left|\mathcal{P} \cap \left(\bigcup _{i=1}^{\alpha} \mathcal{C}_i\right)\right| \nonumber \\ &\qquad \qquad \qquad + |I(\mathcal{P} \cap Z(P), \Gamma_1')| + \sum_{j=2}^{n-1} |I(\mathcal{P} \cap Z(P), \Gamma_j')| + |I(\mathcal{P} \cap Z(P), \Gamma \setminus \Gamma')| \nonumber \\ &\le \frac{C_1}{4}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n+1}} + \left|\mathcal{P} \cap \left(\bigcup _{i=1}^{\alpha} \mathcal{C}_i\right)\right|\sum_{j=2}^{n-1} \theta_j + C_2 \left|\mathcal{P} \cap \left(\bigcup _{i=1}^{\alpha} \mathcal{C}_i\right)\right| \nonumber \\ &\qquad \qquad \qquad + \kappa_4|\mathcal{P} \cap Z(P)| + \sum_{j=2}^{n-1} \theta_j |\mathcal{P} \cap Z(P)| + \kappa_5 |\Gamma| \nonumber \\ &\le \frac{C_1}{4}|\mathcal{P}|^{\frac{s(r+1)}{s(n+r) - n+1} + \varepsilon}|\Gamma|^{\frac{(s-1)(n+r)}{s(n+r)-n+1}} + \left|\mathcal{P}\right|\sum_{j=2}^{n-1} \theta_j + C_2 \left(\left|\mathcal{P}\right| + |\Gamma| \right)\nonumber \end{align*} $$

where the last line holds as long as $C_2 \ge \max \{\kappa _4, \kappa _5\}$ , thus closing the induction.

4.4.2 Proof of Pfaffian joints theorem

Proof of Theorem 4.5

As before, we will use Theorem 3.3 to partition $\{\Gamma _i\}_{i \in [n]}$ ; to distinguish, we let $C_0$ denote the maximum of all constants in Theorem 3.3, and the constant we want to find by C. In general, various intermediary constants that depend on $\varepsilon $ , n, m, $\alpha $ , $\beta $ , and r, shall all be denoted by $\kappa $ without disambiguation; we will use constants $\kappa _i$ only when we need to disambiguate.

We use induction on the total number of elements in $\Gamma _1$ . For the base case, let $|\Gamma _1| = \kappa $ . Consider a particular curve $c_1 \in \Gamma _1$ . With each curve $c_2 \in \Gamma _2$ , by Theorem 2.11 $, c_1$ can have at most $\kappa $ intersections. This means that $c_1$ can have at most $\kappa |\Gamma _2|$ intersections with curves in $\Gamma _2$ . Similarly, $c_1$ can have at most $\kappa |\Gamma _{i}|$ intersections with curves in $\Gamma _i$ for any $2 \le i \le n$ . For a point on $c_1$ to be a joint, the point has to lie on a curve in each $\Gamma _i$ for $2 \le i \le n$ . This in turn means that the number of joints on $c_1$ is bounded by

$$\begin{align*}\kappa\min\{|\Gamma_i|\}_{2 \le i \le n} \le \kappa \left(\prod_{i=2}^n |\Gamma_i|\right)^{\frac{1}{n-1}}.\end{align*}$$

Since $|\Gamma _1| = \kappa $ , we have that the number of joints formed by $\{\Gamma _i\}_{i \in [n]}$ is at most $\kappa _1 \left (\prod _{i=2}^n |\Gamma _i|\right )^{\frac {1}{n-1}}$ ; since $\frac {n+r}{n(n-1)} \ge \frac {1}{n-1}$ , and $\frac {2}{n+1} \ge \frac {1}{n-1}$ whenever $n \ge 3$ , this proves the base case as long as we choose $C \ge \kappa _1$ .

Now suppose that the claim is true for all smaller values of $|\Gamma _1|$ . For the induction step, we use Theorem 3.3 to obtain a $P \in \mathrm {Pfaff}_n(D, \vec {q})$ , with $D = D(\varepsilon , n, m, \alpha , \beta , r)$ to be specified later, such that each connected component of $\mathbb {R}^n \setminus Z(P)$ intersects at most $C_0|\Gamma _i|D^{-(n-1)}$ Pfaffian curves from each $\Gamma _i$ . Suppose $D \ge C_0^{\frac {1}{n-1}} + 1$ . After partitioning, we then have that each connected component has at most $C_0|\Gamma _1|D^{-(n-1)} < |\Gamma _1|$ curves from $\Gamma _1$ , and this means that we can use the induction hypothesis.

Assume without loss of generality that $|\Gamma _1| = \min \left \{|\Gamma _i|\right \}_{i \in [n]}$ , and for brevity, use

$$\begin{align*}\mu := \max\left\{\frac{n + r}{n(n - 1)}, \frac{2}{n+1}\right\}.\end{align*}$$

Using the induction hypothesis, we have that in each connected component of $\mathbb {R}^n \setminus Z(P)$ , there are at most $\mathcal {J}_{cc}$ joints, where

(4.8) $$ \begin{align} \mathcal{J}_{cc} &\le C \cdot \left(C_0|\Gamma_1|D^{-(n-1)}\right)^{\varepsilon} \cdot \prod_{i = 1}^n \left(C_0|\Gamma_i|D^{-(n-1)}\right)^{\mu} \nonumber \\ &= C\cdot C_0^{\varepsilon + n\mu}\cdot \left(|\Gamma_1|D^{-(n-1)}\right)^{\varepsilon} \cdot \prod_{i = 1}^n \left(|\Gamma_i|D^{-(n-1)}\right)^{\mu}\nonumber \\ &= C\cdot C_0^{\varepsilon + n\mu} \cdot \left(D^{-\varepsilon(n-1) - n(n-1)\mu}\right)\cdot |\Gamma_1|^{\varepsilon} \cdot \prod_{i = 1}^n |\Gamma_i|^{\mu} \nonumber \\ &\le C\cdot C_0^{\varepsilon + n\mu} \cdot \left(D^{-\varepsilon(n-1) - (n+r)}\right)\cdot |\Gamma_1|^{\varepsilon} \cdot \prod_{i = 1}^n |\Gamma_i|^{\mu} \qquad\left(\text{ since } \mu \ge \frac{n+r}{n(n-1)}\right). \end{align} $$

By Theorem 1.6 (and also [Reference Gabrielov and Vorobjov25, Theorem 3.4]), we have that $\mathbb {R}^n \setminus Z(P)$ has at most $\kappa D^{n + r}$ connected components, thus the sum total number of joints in all the connected components of $\mathbb {R}^n \setminus Z(P)$ , say $\mathcal {J}_{cc}^{tot}$ , is at most

(4.9) $$ \begin{align} \mathcal{J}_{cc}^{tot} & \le \kappa D^{n+r} \cdot \mathcal{J}_{cc} \nonumber \\ &\le \kappa D^{n+r} \cdot C\cdot C_0^{\varepsilon + n\mu} \cdot \left(D^{-\varepsilon(n-1) - (n+r)}\right)\cdot |\Gamma_1|^{\varepsilon} \cdot \prod_{i = 1}^n |\Gamma_i|^{\mu}\quad (\text{by}\ (4.8)) \nonumber \\&= \kappa_2 \cdot C\cdot C_0^{\varepsilon + n\mu} \cdot D^{-\varepsilon(n-1)}\cdot |\Gamma_1|^{\varepsilon} \cdot \prod_{i = 1}^n |\Gamma_i|^{\mu} \nonumber \\&\le \frac{C}{4} \cdot |\Gamma_1|^{\varepsilon} \cdot \prod_{i = 1}^n |\Gamma_i|^{\mu}, \end{align} $$

where the last inequality is true as long as $D \ge \left (4\kappa _2 \cdot C_0^{\varepsilon + n\mu }\right )^{\frac {1}{\varepsilon (n-1)}}$ .

Having counted the number of joints in $\mathbb {R}^n \setminus Z(P)$ , we shall now count the number of joints on $Z(P)$ . Using Corollary 2.17, we decompose $Z(P)$ into $\kappa $ number of Pfaffian sets, $Z_1, \ldots , Z_{\kappa }$ , such that every joint on $Z(P)$ is on the smooth part of some $Z_i$ . We will fix a $Z_a$ and count the number of joints on the smooth part of $Z_a$ ; the total number of joints on $Z(P)$ will just be at most $\kappa $ times the number of joints on the smooth part of $Z_a$ .

If a joint is a smooth point of $Z_a$ , then it is not possible that all the curves that form the joint are contained in $Z_a$ . This is because if the curves were indeed all contained in $Z_a$ , then the span of their tangent spaces would have been contained in the tangent space of $Z_a$ at the same point, which is at most k-dimensional (by the guarantee of Corollary 2.17), thus contradicting the definition of joint. By Corollary 2.15, we have that a curve that is not contained in $Z_a$ can have at most $\kappa D^{\kappa }$ intersections (the $\kappa $ values could be different, but we don’t disambiguate since it doesn’t affect the final result) with $Z_a$ . This implies that the number of joints for which the curve from $\Gamma _1$ is not contained in $Z_a$ is bounded by $\kappa |\Gamma _1|$ (recall that D is a constant in $\varepsilon , n, m, \alpha , \beta , r$ ). Also, at the same time, for any $2 \le i_1 < i_2 \le n$ , using Corollary 2.15, we also know that the number of joints is bounded by $\kappa |\Gamma _{i_1}||\Gamma _{i_2}|$ ; this is because each curve in $\Gamma _{i_1}$ can intersect each curve in $\Gamma _{i_2}$ in at most $\kappa $ points.

Thus we have that the number of joints on $Z_a^{\text {smooth}}$ for which the curve from $\Gamma _1$ is not contained in $Z_a$ is bounded by

$$ \begin{align*} \min\left\{\kappa |\Gamma_1|, \{\kappa|\Gamma_{i_1}||\Gamma_{i_2}|\}_{2 \le i_1 < i_2 \le n}\right\} &\le \kappa \cdot |\Gamma_1|^{\frac{2}{n+1}} \cdot \prod_{2 \le i_1 < i_2 \le n} \left(|\Gamma_{i_1}||\Gamma_{i_2}|\right)^{\frac{2}{(n+1)(n-2)}} \nonumber \\ &= \kappa \cdot |\Gamma_1|^{\frac{2}{n+1}}\cdot \prod_{i=2}^n \left(|\Gamma_i|^{(n-2)}\right)^{\frac{2}{(n+1)(n-2)}} \nonumber \\ &= \kappa \prod_{i=1}^n |\Gamma_i|^{\frac{2}{n+1}}, \nonumber \end{align*} $$

and consequently, the total number of joints on $Z_a^{\text {smooth}}$ is at most $\kappa \prod _{i=1}^n |\Gamma _i|^{\frac {2}{n+1}}$ . Finally, since $Z(P)$ is composed of $Z_1, \ldots , Z_{\kappa }$ , the total number of joints on $Z(P)$ , say $\mathcal {J}_{Z(P)}$ , is at most

$$\begin{align*}\mathcal{J}_{Z(P)} \le \kappa_3 \cdot \prod_{i=1}^n |\Gamma_i|^{\frac{2}{n+1}} \le \frac{C}{4} \cdot \prod_{i=1}^n |\Gamma_i|^{\frac{2}{n+1}} \le \frac{C}{4} \cdot \prod_{i=1}^n |\Gamma_i|^{\mu}, \end{align*}$$

where the second inequality follows as long as $C \ge 4\kappa _3$ .

Finally, we get that the total number of joints is at most

$$\begin{align*}\mathcal{J}_{cc}^{tot} + \mathcal{J}_{Z(P)} \le \frac{C}{4} \cdot |\Gamma_1|^{\varepsilon} \cdot \prod_{i=1}^n |\Gamma_i|^{\mu} + \frac{C}{4} \cdot \prod_{i=1}^n |\Gamma_i|^{\mu} \le C \cdot |\Gamma_1|^{\varepsilon} \cdot \prod_{i=1}^n |\Gamma_i|^{\mu}.\end{align*}$$

Thus as long as we choose

$$\begin{align*}D \ge \max\left\{C_0^{\frac{1}{n-1}} + 1, \left(4\kappa_2 \cdot C_0^{\varepsilon + n\mu}\right)^{\frac{1}{\varepsilon(n-1)}}\right\} \text{ and } C \ge \max\left\{\kappa_1, 4\kappa_3\right\}, \end{align*}$$

we can close the induction.

A Proof of Theorem 3.2

Proof of Theorem 3.2

Recall that $\mathbb {R}[X_1, \ldots , X_n]_{(\le d)}$ has dimension $\binom {d + n}{n}> \frac {d^n}{n!}$ . For $\ell \in \mathbb {N}_{> 0}$ , define $d_{\ell }:= \left \lceil \left ( m\cdot n!\cdot 2^{\ell -1}\right )^{1/n}\right \rceil $ , so that

(A.1) $$ \begin{align} m2^{\ell-1} \le \frac{d_{\ell}^n}{n!} < m2^{n + \ell - 1}. \end{align} $$

Let s be the smallest integer such that $\sum _{\ell =1}^s d_{\ell } \le D < \sum _{\ell =1}^{s+1} d_{\ell }$ . Note that we do not exclude $s=0$ . In particular,

(A.2)

where $\alpha _1 = \alpha _1(n)$ is a constant.

By (A.1), we have

$$\begin{align*}\mathrm{dim}\, \mathbb{R}[X_1, \ldots, X_n]_{(\le d_{\ell})}> \frac{d_{\ell}^n}{n!} \ge m 2^{\ell-1} \end{align*}$$

for every $\ell $ . Let $V_\ell $ be an arbitrary subspace of $\mathbb {R}[X_1, \ldots , X_n]_{(\le d_{\ell })}$ of dimension exactly $m2^{\ell -1} + 1$ , and let $S(V_\ell )$ denote the unit sphere in $V_\ell $ . Define

$$ \begin{align*} Y := \prod_{\ell=1}^s S(V_\ell) \cong \prod_{\ell=1}^s S^{m2^{\ell-1}}. \end{align*} $$

For any polynomial $P \in \mathbb {R}[X_1, \ldots , X_n]$ , let $C_P^{+} \subseteq \mathbb {R}^n$ denote the set of points in $\mathbb {R}^n$ where P is positive, and let $C_P^{-}$ be the set of points where P is negative, so that $\mathbb {R}^n \setminus Z(P) = C_P^{+} \cup C_P^{-}$ and $C_P^{-} \cap C_P^{+} = \emptyset $ . Let $(P_1, \ldots , P_s) \in Y$ be a tuple of polynomials and set $P=P_1\cdots P_s$ . It is clear from (A.2) that $\mathrm {deg} \, P \le \sum _{\ell =1}^s d_{\ell } \le D$ . For $\sigma = (\sigma _1, \ldots , \sigma _s) \in \mathbb {Z}_2^{s}$ , we define the sign condition

$$ \begin{align*} \mathcal{O}_{\sigma}^{(P_1,\dots,P_s)} := \bigcap_{i=1}^s C_{P_i}^{\sigma_i}. \end{align*} $$

This implies $\mathbb {R}^n \setminus Z(P) = \bigcup _{\sigma \in \mathbb {Z}_2^s} \mathcal {O}_{\sigma }^{(P_1,\ldots , P_s)}$ and each $\mathcal {O}_{\sigma }^{(P_1, \ldots , P_s)}$ is a union of some connected components of $\mathbb {R}^n \setminus Z(P)$ .

For any $\sigma \in \mathbb {Z}_2^s$ and semi-Pfaffian set $\mathcal {X} \subseteq \mathbb {R}^n$ , define the indicator function $I_{\sigma , \mathcal {X}}: Y \rightarrow \mathbb {R}$ ,

$$ \begin{align*} I_{\sigma, \mathcal{X}}(P_1, \ldots, P_s)=\begin{cases} 1, & \text{if } \mathcal{O}_{\sigma}^{(P_1, \ldots, P_s)} \cap \mathcal{X} \neq \emptyset,\\ 0, & \text{if } \mathcal{O}_{\sigma}^{(P_1, \ldots, P_s)} \cap \mathcal{X} = \emptyset \end{cases}. \end{align*} $$

Also, let $U_s \subseteq \mathbb {R}^{2^s}$ be the codimension 1 subspace defined as

(A.3) $$ \begin{align} U_s = \left\{(y_{\sigma})_{\sigma \in \mathbb{Z}_2^s} \in R^{2^s} \boldsymbol{:} \sum_{\sigma \in \mathbb{Z}_2^s} y_{\sigma} = 0\right\}. \end{align} $$

Consider the following map $\Phi : Y \rightarrow \bigoplus _{i=1}^m U_s$ , which takes

(A.4) $$ \begin{align} (P_1, \ldots, P_s) \mapsto \left(\left(\sum_{\mathcal{X} \in \Gamma_i} I_{\sigma, \mathcal{X}}(P_1, \ldots, P_s) - \frac{1}{2^s} \sum_{\beta \in \mathbb{Z}_2^s} \sum_{\mathcal{X} \in \Gamma_i} I_{\beta, \mathcal{X}}(P_1, \ldots, P_s)\right)_{\sigma \in \mathbb{Z}_2^s}\right)_{i \in [m]}. \end{align} $$

The sum $\sum _{\mathcal {X} \in \Gamma _i} I_{\sigma , \mathcal {X}}(P_1, \ldots , P_s)$ counts the number of varieties in $\Gamma _i$ that intersect the sign condition $\mathcal {O}_{\sigma }^{(P_1, \ldots , P_s)}$ . Now, assume that the map $\Phi $ has a zero and pick $(P_1, \ldots , P_s) \in \Phi ^{-1}(0)$ . This means that each sign condition in $\{\mathcal {O}_{\sigma }^{(P_1, \ldots , P_s)}\}_{\sigma \in \mathbb {Z}_2^s}$ is intersected by the same number of elements of the set $\Gamma _i$ , for all $i \in [m]$ .

For all $i \in [m]$ with $k_i> 0$ , by Theorem 1.6, we have that each $\mathcal {X} \in \Gamma _i$ intersects at most $CD^{k_i + r_i}$ connected components of $\mathbb {R}^n \setminus Z(P)$ , and since each sign condition is a disjoint union of connected components of $\mathbb {R}^n \setminus Z(P)$ , we have that each $\mathcal {X} \in \Gamma _i$ intersects at most $CD^{k_i + r_i}$ sign conditions. Since we have that $(P_1, \ldots , P_s) \in \Phi ^{-1}(0)$ , we have that, for all $i \in [m]$ , each sign condition intersects at most $\frac {1}{2^s}|\Gamma _i|CD^{k_i + r_i} < C'|\Gamma _i|mD^{k_i + r_i - n}$ semi-Pfaffian sets from $\Gamma _i$ , where we have used (A.2) to obtain the upper bound. This proves Part 3.2.

On the other hand, for all $i \in [m]$ with $k_i = 0$ , each $\mathcal {X} \in \Gamma _i$ intersects at most C connected components of $\mathbb {R}^n \setminus Z(P)$ . In this specific case, we get that each connected component of $\mathbb {R}^n \setminus Z(P)$ intersects at most $C'|\Gamma _i|mD^{-n}$ elements of $\Gamma _i$ . This proves Part 3.2.

It remains to prove that $\Phi $ has a zero. Towards this, since the indicator functions $I_{\sigma , \mathcal {X}}$ are not continuous, we have to consider continuous approximations of the indicators. In [Reference Guth28] the following lemma is proved which allows us to approximate $I_{\sigma , \mathcal {X}}$ by sequences of continuous functions.

Lemma A.1 (Lemma 3.1 in [Reference Guth28])

For every $\varepsilon> 0$ , $\mathcal {X} \subseteq \mathbb {R}^n$ , and $\sigma \in \mathbb {Z}_2^s$ , there exist functions $I_{\sigma , \mathcal {X}}^{\varepsilon }: Y \rightarrow \mathbb {R}$ with the following properties.

  1. 1. $I_{\sigma , \mathcal {X}}^{\varepsilon }$ are continuous.

  2. 2. $0 \le I_{\sigma , \mathcal {X}}^{\varepsilon } \le 1$ .

  3. 3. If $\mathcal {O}_{\sigma }^{(P_1, \ldots , P_s)} \cap \mathcal {X} = \emptyset $ , then $I_{\sigma , \mathcal {X}}^{\varepsilon } = 0$ .

  4. 4. If $\varepsilon _i \rightarrow 0$ , $(P_1^i, \ldots , P_s^i) \rightarrow (P_1, \ldots , P_s)$ in Y and $\mathcal {O}_{\sigma }^{(P_1, \ldots , P_s)} \cap \mathcal {X} \neq \emptyset $ , then

    $$\begin{align*}\lim_{i \rightarrow \infty} I_{\sigma, \mathcal{X}}^{\varepsilon_i}(P_1^i, \ldots, P_s^i) = 1.\end{align*}$$
    In other words, $I_{\sigma , \mathcal {X}}(P_1, \ldots , P_s) \le \lim \inf _{i \rightarrow \infty } I_{\sigma , \mathcal {X}}^{\varepsilon _i}(P_1^i, \ldots , P_s^i)$ .

For some sequence $\varepsilon _i \rightarrow 0$ , let $\Phi _i: Y \rightarrow U_s^{\oplus j}$ be defined as

$$\begin{align*}(P_1, \ldots, P_s) \mapsto \left(\left(\sum_{\mathcal{X} \in \Gamma_i} I^{\varepsilon_i}_{\sigma, \mathcal{X}}(P_1, \ldots, P_s) - \frac{1}{2^s} \sum_{\beta \in \mathbb{Z}_2^s} \sum_{\mathcal{X} \in \Gamma_i} I^{\varepsilon_i}_{\beta, \mathcal{X}}(P_1, \ldots, P_s)\right)_{\sigma \in \mathbb{Z}_2^s}\right)_{i \in [m]}. \end{align*}$$

Consider the following action of $\mathbb {Z}_2^s = \langle \omega _1, \ldots , \omega _s\rangle $ on Y: for all $1 \le l \le s$ , given $(P_1, \ldots , P_s) \in Y$ , the generator $\omega _l$ acts as

(A.5) $$ \begin{align} \omega_l \circ (P_1, \ldots, P_s) = (P_1, \ldots, -P_l, \ldots, P_s). \end{align} $$

Also, consider the following action of $\mathbb {Z}^s_2$ on $\mathbb {R}^{2^s}$ : an element $(\beta _1, \ldots , \beta _s) \in \mathbb {Z}_2^s$ acts on $(y_{\alpha })_{\alpha \in \mathbb {Z}_2^s}$ by acting on the indices as

(A.6) $$ \begin{align} (\beta_1, \ldots, \beta_s) \circ (\alpha_1, \ldots, \alpha_s) = (\beta_1 + \alpha_1, \ldots, \alpha_s + \beta_s), \end{align} $$

where the addition is in $\mathbb {Z}_2$ . Thus the vector subspace $U_s$ is a $\mathbb {Z}_2^s$ -subrepresentation of $\mathbb {R}^{2^s}$ .

The functions $\Phi _i$ are continuous and are $\mathbb {Z}_2^s$ -equivariant with respect to the above actions ((A.5) and (A.6)), assuming the diagonal action on the direct sum $U_s^{\oplus j}$ . By the use of the $\mathbb {Z}_2^s$ -equivariance of $\Phi _i$ and some tools from equivariant cohomology (Fadell-Husseini index theory [Reference Fadell and Husseini22]), it is proved in [Reference Blagojević, Blagojević and Ziegler13, Section 2.7] that each $\Phi _i$ indeed has a zero.

Since each $\Phi _i$ has a zero, this means that for each i, there exists a $\rho _i \in Y$ such that $\Phi _i(\rho _i) = 0$ . Because Y is compact, this means that there is a converging subsequence

$$\begin{align*}\lim_{i \rightarrow \infty} \rho_{\kappa(i)} = \rho \in Y, \end{align*}$$

where $\kappa : \mathbb {N} \rightarrow \mathbb {N}$ is a strictly increasing function. Since by Lemma A.1 we have that the indicator functions $I_{\sigma , \mathcal {X}}$ are limits of sequences of continuous maps, we have that

$$\begin{align*}\Phi(\rho) = \lim_{i \rightarrow \infty} \Phi_{\kappa(i)}(\rho_{\kappa(i)}) = 0. \end{align*}$$

This proves that $\Phi $ has a zero, and completes the proof.

Acknowledgments

AN would like to especially thank Adam Sheffer for his comments and very helpful discussions. AN would also like to thank Saugata Basu and Joshua Zahl for their comments on a preliminary draft, and Alexander Balsera for useful discussions.

Competing interests

The authors have no competing interest to declare.

Financial support

AN was supported by EPSRC Grant EP/V003542/1.

Footnotes

1 The subscript $\mathcal {X}$ in the big-O notation is shorthand to denote that we are considering various parameters that determine the complexity of $\mathcal {X}$ to be fixed; our theorem statements are more precise.

References

Afshani, P. and Cheng, P., ‘Lower bounds for semialgebraic range searching and stabbing problems’, J. ACM, 70(2) (2023), 126.10.1145/3578574CrossRefGoogle Scholar
Agarwal, P. K., Aronov, B., Ezra, E. and Zahl, J., ‘Efficient algorithm for generalized polynomial partitioning and its applications’, SIAM J. Comput. 50(2) (2021), 760787.10.1137/19M1268550CrossRefGoogle Scholar
Anderson, A., ‘Combinatorial bounds in distal structures’, J. Symb. Log. (2023), 133.10.1017/jsl.2023.74CrossRefGoogle Scholar
Aronov, B., Ezra, E. and Zahl, J., ‘Constructive polynomial partitioning for algebraic curves in ${\mathbb{R}}^3$ with applications’, SIAM J. Comput. 49(6) (2020), 11091127.10.1137/19M1257548CrossRefGoogle Scholar
Balsera, A., ‘Incidences with Pfaffian curves and functions’, arXiv preprint arXiv:2311.05517, 2023.Google Scholar
Barone, S. and Basu, S., ‘On a real analog of Bezout inequality and the number of connected components of sign conditions’, Proc. Lond. Math. Soc. 112(1) (2016), 115145.10.1112/plms/pdv059CrossRefGoogle Scholar
Basit, A., Chernikov, A., Starchenko, S., Tao, T. and Tran, C-M., ‘Zarankiewicz’s problem for semilinear hypergraphs’, Forum Math. Sigma 9 (2021), e59.10.1017/fms.2021.52CrossRefGoogle Scholar
Basu, S., Lerario, A. and Natarajan, A., ‘Zeroes of polynomials on definable hypersurfaces: pathologies exist, but they are rare’, Q. J. Math. 70(4) (2019), 13971409.10.1093/qmath/haz022CrossRefGoogle Scholar
Basu, S. and Raz, O. E, ‘An o-minimal Szemerédi–Trotter theorem’, Q. J. Math. 69(1) (2018), 223239.10.1093/qmath/hax037CrossRefGoogle Scholar
Basu, S. and Sombra, M., ‘Polynomial partitioning on varieties of codimension two and point-hypersurface incidences in four dimensions’, Discrete Comput. Geom. 55(1) (2016), 158184.10.1007/s00454-015-9736-4CrossRefGoogle Scholar
Bennett, J., Carbery, A. and Tao, T., ‘On the multilinear restriction and Kakeya conjectures’, Acta Math. 196(2) (2006), 261302.10.1007/s11511-006-0006-4CrossRefGoogle Scholar
Binyamini, G. and Novikov, D., ‘Tameness in geometry and arithmetic: beyond o-minimality’, in D. Beliaev and S. Smirnov (eds.), ICM – International Congress of Mathematicians. Vol. 3. Sections 1–4 (EMS Press, Berlin, 2023), 14401461.10.4171/icm2022/117CrossRefGoogle Scholar
Blagojević, P. V. M., Blagojević, A. S. D. and Ziegler, G. M., ‘Polynomial partitioning for several sets of varieties’, J. Fixed Point Theory Appl. 19 (2017), 16531660.10.1007/s11784-016-0322-zCrossRefGoogle Scholar
Chazelle, B., Edelsbrunner, H., Guibas, L. J, Pollack, R., Seidel, R., Sharir, M. and Snoeyink, J., ‘Counting and cutting cycles of lines and rods in space’, Comput. Geom. 1(6) (1992), 305323.10.1016/0925-7721(92)90009-HCrossRefGoogle Scholar
Chernikov, A., Galvin, D. and Starchenko, S., ‘Cutting lemma and Zarankiewicz’s problem in distal structures’, Sel. Math. 26 (2020), 127.10.1007/s00029-020-0551-2CrossRefGoogle Scholar
Chernikov, A., Peterzil, Y. and Starchenko, S., ‘Model-theoretic Elekes-Szabó for stable and o-minimal hypergraphs’, Duke Math. J. 173(3) (2024), 419512.10.1215/00127094-2023-0018CrossRefGoogle Scholar
Chernikov, A. and Starchenko, S., ‘Regularity lemma for distal structures’, J. Eur. Math. Soc. (JEMS) 20(10) (2018), 24372466.10.4171/jems/816CrossRefGoogle Scholar
Chernikov, A., Starchenko, S. and Thomas, M. E. M., ‘Ramsey growth in some nip structures’, J. Inst. Math. Jussieu 20(1) (2021), 129.10.1017/S1474748019000100CrossRefGoogle Scholar
Clarkson, K. L., Edelsbrunner, H., Guibas, L. J., Sharir, M. and Welzl, E., ‘Combinatorial complexity bounds for arrangements of curves and spheres’, Discrete Comput. Geom. 5(2) (1990), 99160.10.1007/BF02187783CrossRefGoogle Scholar
Dvir, Z., ‘Incidence theorems and their applications’, Found. Trends Theor. Comput. Sci. 6(4) (2010), 257393.10.1561/0400000056CrossRefGoogle Scholar
Elekes, G., ‘On the number of sums and products’, Acta Arith. 81(4) (1997), 365367.10.4064/aa-81-4-365-367CrossRefGoogle Scholar
Fadell, E. and Husseini, S., ‘An ideal-valued cohomological index theory with applications to Borsuk-Ulam and Bourgin-Yang theorems’, Ergod. Theory Dyn. Syst. 8 ${}^{\ast }$ (1988), 7385.Google Scholar
Feldman, S. and Sharir, M., ‘An improved bound for joints in arrangements of lines in space’, Discrete Comput. Geom. 33 (2005), 307320.10.1007/s00454-004-1093-7CrossRefGoogle Scholar
Fox, J., Pach, J., Sheffer, A., Suk, A. and Zahl, J., ‘A semi-algebraic version of Zarankiewicz’s problem’, J. Eur. Math. Soc. (JEMS) 19(6) (2017), 17851810.10.4171/jems/705CrossRefGoogle Scholar
Gabrielov, A. and Vorobjov, N., ‘Complexity of computations with Pfaffian and Noetherian functions’, Normal forms, bifurcations and finiteness problems in differential equations, vol. 137 NATO Sci. Ser. II Math. Phys. Chem. (Kluwer Acad. Publ., Dordrecht, 2004), 211250.10.1007/978-94-007-1025-2_5CrossRefGoogle Scholar
Gabrielov, A. and Vorobjov, N., ‘Approximation of definable sets by compact families, and upper bounds on homotopy and homology’, J. Lond. Math. Soc. (2), 80(1) (2009), 3554.10.1112/jlms/jdp006CrossRefGoogle Scholar
Grothendieck, A., ‘Esquisse d’un programme’, in Geometric Galois actions, 1, vol. 242 of London Math. Soc. Lecture Note Ser. (Cambridge Univ. Press, Cambridge, 1997), 548. With an English translation on pp. 243–283.Google Scholar
Guth, L., ‘Polynomial partitioning for a set of varieties’, in Mathematical Proceedings of the Cambridge Philosophical Society, vol. 159 (Cambridge University Press, 2015), 459469.Google Scholar
Guth, L., Polynomial Methods in Combinatorics, vol. 64 (American Mathematical Society, Providence, RI, 2016), ix+273.10.1090/ulect/064CrossRefGoogle Scholar
Guth, L., ‘A restriction estimate using polynomial partitioning’, J. Am. Math. Soc. 29(2) (2016), 371413.10.1090/jams827CrossRefGoogle Scholar
Guth, L., ‘Restriction estimates using polynomial partitioning II’, Acta Math. 221(1) (2018), 81142.10.4310/ACTA.2018.v221.n1.a3CrossRefGoogle Scholar
Guth, L. and Katz, N. H., ‘Algebraic methods in discrete analogs of the Kakeya problem’, Adv. Math. 225(5) (2010), 28282839.10.1016/j.aim.2010.05.015CrossRefGoogle Scholar
Guth, L. and Katz, N. H., ‘On the Erdős distinct distances problem in the plane’, Annals of Mathematics. Second Series 181(1) (2015), 155190.10.4007/annals.2015.181.1.2CrossRefGoogle Scholar
Gwoździewicz, J., Kurdyka, K. and Parusiński, A., ‘On the number of solutions of an algebraic equation on the curve $y={e}^x+\sin x$ , $x>0$ , and a consequence for o-minimal structures’, Proc. Am. Math. Soc. 127(4) (1999), 10571064.10.1090/S0002-9939-99-04672-9CrossRef0$+,+and+a+consequence+for+o-minimal+structures’,+Proc.+Am.+Math.+Soc.+127(4)+(1999),+1057–1064.>Google Scholar
Ilyashenko, Yu., ‘Centennial history of Hilbert’s 16th problem’, Bull. Amer. Math. Soc. (N.S.) 39(3) (2002), 301354.10.1090/S0273-0979-02-00946-1CrossRefGoogle Scholar
Kaplan, H., Matoušek, J. and Sharir, M., ‘Simple proofs of classical theorems in discrete geometry via the Guth–Katz polynomial partitioning technique’, Discrete Comput. Geom. 48 (2012), 499517.10.1007/s00454-012-9443-3CrossRefGoogle Scholar
Kaplan, H., Sharir, M. and Shustin, E., ‘On lines and joints’, Discrete Comput. Geom. 44(4) (2010), 838843.10.1007/s00454-010-9246-3CrossRefGoogle Scholar
Karpinski, M. and Macintyre, A., ‘Polynomial bounds for VC dimension of sigmoidal and general Pfaffian neural networks’, J. Comput. Syst. Sci. 54(1) (1997), 169176.10.1006/jcss.1997.1477CrossRefGoogle Scholar
Khovanskiĭ, A. G., ‘A class of systems of transcendental equations’, in Doklady Akademii Nauk, vol. 255 (Russian Academy of Sciences, Russia, 1980), 804807.Google Scholar
Khovanskiĭ, A. G., ‘Fewnomials and Pfaff manifolds’, in Proceedings of the International Congress of Mathematicians, Vols. 1, 2 (Warsaw, 1983) (PWN, Warsaw, 1984), 549564.Google Scholar
Khovanskiĭ, A. G., Fewnomials, vol. 88 Translations of Mathematical Monographs, (American Mathematical Society, Providence, RI, 1991), viii+139.10.1090/mmono/088CrossRefGoogle Scholar
Krantz, S. G. and Parks, H. R., A Primer of Real Analytic Functions. Birkhäuser Advanced Texts: Basler Lehrbücher [Birkhäuser Advanced Texts: Basel Textbooks], second edition (Birkhäuser Boston, Inc., Boston, MA, 2002).10.1007/978-0-8176-8134-0CrossRefGoogle Scholar
Lion, J.-M. and Rolin, J-P., ‘Volumes, feuilles de Rolle de feuilletages analytiques et théorème de Wilkie’, Ann. Fac. Sci. Toulouse Math. (6), 7(1) (1998), 93112.10.5802/afst.891CrossRefGoogle Scholar
Marker, D., Model Theory: An Introduction, vol. 217 of Graduate Texts in Mathematics (Springer-Verlag, New York, 2002). An introduction.Google Scholar
Matoušek, J., Lectures on Discrete Geometry, vol. 212 of Graduate Texts in Mathematics (Springer-Verlag, New York, 2002).10.1007/978-1-4613-0039-7CrossRefGoogle Scholar
Matoušek, J. and Patáková, Z., ‘Multilevel polynomial partitions and simplified range searching’, Discrete Comput. Geom. 54(1) (2015), 2241.10.1007/s00454-015-9701-2CrossRefGoogle Scholar
Miller, C. and Speissegger, P., ‘Pfaffian differential equations over exponential o-minimal structures’, J. Symbolic Logic 67(1) (2002), 438448.10.2178/jsl/1190150053CrossRefGoogle Scholar
Milnor, J., ‘On the Betti numbers of real varieties’, Proc. Am. Math. Soc. 15(2) (1964), 275280.10.1090/S0002-9939-1964-0161339-9CrossRefGoogle Scholar
Natarajan, A., Betti numbers of deterministic and random sets in semi-algebraic and o-minimal geometry. PhD Thesis. Purdue University, West Lafayette, Indiana, USA, 2020.Google Scholar
Osgood, W. F., ‘On functions of several complex variables’, Trans. Amer. Math. Soc. 17(1) (1916), 18.10.1090/S0002-9947-1916-1501027-1CrossRefGoogle Scholar
Pach, J. and Sharir, M., ‘On the number of incidences between points and curves’, Combin. Probab. Comput. 7(1) (1998), 121127.10.1017/S0963548397003192CrossRefGoogle Scholar
Petrovskii, I. G. and Oleinik, O. A., ‘On the topology of real algebraic surfaces’, Izvestiya Rossiiskoi Akademii Nauk. Seriya Matematicheskaya 13(5) (1949), 389402.Google Scholar
Quilodrán, R., ‘The joints problem in ${\mathbb{R}}^n$ ’, SIAM J. Discrete Math. 23(4) (2010), 22112213.10.1137/090763160CrossRefGoogle Scholar
Roy, M-F. and Vorobjov, N., ‘Finding irreducible components of some real transcendental varieties’, Comput. Complex. 4 (1994), 107132.10.1007/BF01202285CrossRefGoogle Scholar
Scanlon, T., ‘O-minimality’, Gazette des Mathématiciens (149) (2016), 3339.Google Scholar
Sharir, M., ‘On joints in arrangements of lines in space and related problems’, J. Comb. Theory Ser. A 67(1) (1994), 8999.10.1016/0097-3165(94)90005-1CrossRefGoogle Scholar
Sharir, M., Sheffer, A. and Solomon, N., ‘Incidences with curves in ${\mathbb{R}}^d$ ’, Electron. J. Combin. 23(4) (2016), Paper 4.16, 16.10.37236/5840CrossRefGoogle Scholar
Sheffer, A., Polynomial Methods and Incidence Theory, vol. 197 of Cambridge Studies in Advanced Mathematics (Cambridge University Press, Cambridge, 2022).10.1017/9781108959988CrossRefGoogle Scholar
Sheffer, A. and Zahl, J., ‘Distinct distances in the complex plane’, Trans. Am. Math. Soc. 374(09) (2021), 66916725.10.1090/tran/8420CrossRefGoogle Scholar
Solymosi, J. and Tao, T., ‘An incidence theorem in higher dimensions’, Discrete Comput. Geom. 48 (2012), 255280.10.1007/s00454-012-9420-xCrossRefGoogle Scholar
Speissegger, P., ‘The Pfaffian closure of an o-minimal structure’, J. Reine Angew. Math. 508 (1999), 189211.10.1515/crll.1999.508.189CrossRefGoogle Scholar
Szemerédi, E. and Trotter, W. T. Jr., ‘Extremal problems in discrete geometry’, Combinatorica 3(3–4) (1983), 381392.10.1007/BF02579194CrossRefGoogle Scholar
Thom, R., ‘Sur l’homologie des variétés algébriques réelles’, in Differential and Combinatorial Topology (A Symposium in Honor of Marston Morse) (Princeton Univ. Press, Princeton, NJ, 1965), 255265.10.1515/9781400874842-016CrossRefGoogle Scholar
Tidor, J., Yu, H-H. H. and Zhao, Y., ‘Joints of varieties’, Geom. Funct. Anal. 32(2) (2022), 302339.10.1007/s00039-022-00597-5CrossRefGoogle Scholar
den Dries, L. V., Tame Topology and O-Minimal Structures, vol. 248 (Cambridge University Press, Cambridge, 1998).10.1017/CBO9780511525919CrossRefGoogle Scholar
Walsh, M. N., ‘The polynomial method over varieties’, Invent. Math. 222(2) (2020), 469512.10.1007/s00222-020-00975-6CrossRefGoogle Scholar
Wilkie, A., ‘Model completeness results for expansions of the ordered field of real numbers by restricted Pfaffian functions and the exponential function’, J. Am. Math. Soc. 9(4) (1996), 10511094.10.1090/S0894-0347-96-00216-0CrossRefGoogle Scholar
Wolff, T., ‘Recent work connected with the Kakeya problem’, in Prospects in Mathematics (Princeton, NJ, 1996) (American Mathematical Society, Providence, RI, 1999), 129162.Google Scholar
Yang, B., ‘Generalizations of joints problem’, 2016, arXiv preprint arXiv:1606.08525.Google Scholar