This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Improved effective Łojasiewicz inequality and applications

Saugata Basu Department of Mathematics, Purdue University, West Lafayette, IN 47906, U.S.A. [email protected]  and  Ali Mohammad-Nezhad Department of Mathematical Sciences, Carnegie Mellon University, Pittsburgh, PA 15123, U.S.A. [email protected]
Abstract.

Let R\mathrm{R} be a real closed field. Given a closed and bounded semi-algebraic set ARnA\subset\mathrm{R}^{n} and semi-algebraic continuous functions f,g:ARf,g:A\rightarrow\mathrm{R}, such that f1(0)g1(0)f^{-1}(0)\subset g^{-1}(0), there exist NN and cRc\in\mathrm{R}, such that the inequality (Łojasiewicz inequality) |g(x)|Nc|f(x)||g(x)|^{N}\leq c\cdot|f(x)| holds for all xAx\in A. In this paper we consider the case when AA is defined by a quantifier-free formula with atoms of the form P=0,P>0,P𝒫P=0,P>0,P\in\mathcal{P} for some finite subset of polynomials 𝒫R[X1,,Xn]d\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d}, and the graphs of f,gf,g are also defined by quantifier-free formulas with atoms of the form Q=0,Q>0,Q𝒬Q=0,Q>0,Q\in\mathcal{Q}, for some finite set 𝒬R[X1,,Xn,Y]d\mathcal{Q}\subset\mathrm{R}[X_{1},\ldots,X_{n},Y]_{\leq d}. We prove that the Łojasiewicz exponent NN in this case is bounded by (8d)2(n+7)(8d)^{2(n+7)}. Our bound depends on dd and nn, but is independent of the combinatorial parameters, namely the cardinalities of 𝒫\mathcal{P} and 𝒬\mathcal{Q}. The previous best known upper bound in this generality appeared in P. Solernó, Effective Łojasiewicz Inequalities in Semi-algebraic Geometry, Applicable Algebra in Engineering, Communication and Computing (1991) and depended on the sum of degrees of the polynomials defining A,f,gA,f,g and thus implicitly on the cardinalities of 𝒫\mathcal{P} and 𝒬\mathcal{Q}. As a consequence we improve the current best error bounds for polynomial systems under some conditions. Finally, as an abstraction of the notion of independence of the Łojasiewicz exponent from the combinatorial parameters occurring in the descriptions of the given pair of functions, we prove a version of Łojasiewicz inequality in polynomially bounded o-minimal structures. We prove the existence of a common Łojasiewicz exponent for certain combinatorially defined infinite (but not necessarily definable) families of pairs of functions. This improves a prior result of Chris Miller (C. Miller, Expansions of the real field with power functions, Ann. Pure Appl. Logic (1994)).

Key words and phrases:
Łojasiewicz inequality, combinatorial complexity, error bounds, polynomially bounded o-minimal structures
2000 Mathematics Subject Classification:
Primary 14P10; Secondary 03C64, 90C23

1. Introduction

L. Schwartz conjectured that if ff is a real analytic function and TT a distribution in some open subset ΩRn\Omega\subset\mathrm{R}^{n}, then there exists a distribution SS satisfying fS=TfS=T. As a main tool in proving this conjecture, Łojasiewicz [31] proved that if VV is the set of real zeros of ff, and xx in a sufficiently small neighborhood of a point x0x_{0} in VV, there exists a constant dd such that

|f(x)|ddist(x,V)d,|f(x)|\geq d\cdot{\rm dist}(x,V)^{d},

where dist(x,V){\rm dist}(x,V) denotes the distance of xx from VV. In case ff is a polynomial the result was obtained by Hörmander [18].

Several variants of Łojasiewicz inequality have appeared in the literature both in the semi-algebraic and analytic categories. In the semi-algebraic category, the following slightly more general version of the above inequality appears in [10, Corollary 2.6.7].

Unless otherwise specified R\mathrm{R} is a fixed real closed field for the rest of the paper.

Let ARnA\subset\mathrm{R}^{n} be a closed and bounded semi-algebraic set and let f,g:ARf,g:A\to\mathrm{R} be continuous semi-algebraic functions. Furthermore, suppose that f1(0)g1(0)f^{-1}(0)\subset g^{-1}(0). Then there exist  cRc\in\mathrm{R} and a rational number ρ\rho depending on AA, ff, and gg, such that

(1.1) |g(x)|ρc|f(x)|,xA.|g(x)|^{\rho}\leq c\cdot|f(x)|,\quad\forall x\in A.

We denote the infimum of ρ\rho by (f,gA)\mathcal{L}(f,g\mid A) which is called the Łojasiewicz exponent.

The inequality (1.1) is usually called the Łojasiewicz inequality and has found many applications (independent of the division problem of L. Schwartz) – for example, in singularity theory, partial differential equations, and optimization. We survey some of these applications later in the paper and improve some of these results using the version of Łojasiewicz inequality proved in the current paper.

Driven by the applications mentioned above there has been a lot of interest in obtaining effective bounds on (f,gA)\mathcal{L}(f,g\mid A).

2. Main results

In this paper we prove new quantitative versions of the inequality (1.1) in the semi-algebraic (and more generally in the o-minimal context). Before stating our results we introduce a few necessary definitions.

Definition 2.1 (𝒫\mathcal{P}-formulas and semi-algebraic sets).

Let 𝒫R[X1,,Xn]\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}], 𝒬R[X1,,Xn,Y]\mathcal{Q}\subset\mathrm{R}[X_{1},\ldots,X_{n},Y] be a finite sets of polynomials. We will call a quantifier-free first-order formula (in the theory of the reals) with atoms P=0,P>0,P<0,P𝒫P=0,P>0,P<0,P\in\mathcal{P} to be a 𝒫\mathcal{P}-formula. Given any first-order formula Φ(X1,,Xn)\Phi(X_{1},\ldots,X_{n}) in the theory of the reals (possibly with quantifiers), we will denote by (Φ,Rn){\mathcal{R}}(\Phi,\mathrm{R}^{n}) the set of points of Rn\mathrm{R}^{n} satisfying Φ\Phi, and call (Φ,Rn){\mathcal{R}}(\Phi,\mathrm{R}^{n}) the realization of Φ\Phi. We will call the realization of a 𝒫\mathcal{P}-formula a 𝒫\mathcal{P}-semi-algebraic set. A 𝒬\mathcal{Q}-semi-algebraic function is a function whose graph is a 𝒬\mathcal{Q}-semi-algebraic set.

We denote by R[X1,,Xn]d\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d} the subset of polynomials in R[X1,,Xn]\mathrm{R}[X_{1},\ldots,X_{n}] with degrees d\leq d.

2.1. Semi-algebraic case

We prove the following theorem in the semi-algebraic setting which improves the currently the best known upper bound [45] in a significant way (see Section 4 below).

Theorem 1.

Let d2d\geq 2, 𝒫R[X1,,Xn]d,𝒬R[X1,,Xn,Y]d\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d},\mathcal{Q}\subset\mathrm{R}[X_{1},\ldots,X_{n},Y]_{\leq d}, ARnA\subset\mathrm{R}^{n} a closed and bounded 𝒫\mathcal{P}-semi-algebraic set, and f,g:ARf,g:A\rightarrow\mathrm{R} continuous 𝒬\mathcal{Q}-semi-algebraic functions, satisfying f1(0)g1(0)f^{-1}(0)\subset g^{-1}(0).

Then there exist c=c(A,f,g)Rc=c(A,f,g)\in\mathrm{R} and N(8d)2(n+7)N\leq(8d)^{2(n+7)} such that for all xAx\in A,

(2.1) |g(x)|Nc|f(x)|.|g(x)|^{N}\leq c\cdot|f(x)|.

In other words,

(2.2) (f,gA)(8d)2(n+7)=dO(n).\mathcal{L}(f,g\mid A)\leq(8d)^{2(n+7)}=d^{O(n)}.

In the special case where R=\mathrm{R}=\mathbb{R} and

𝒫[X1,,Xn]d,𝒬[X1,,Xn,Y]d,\mathcal{P}\subset\mathbb{Z}[X_{1},\ldots,X_{n}]_{\leq d},\mathcal{Q}\subset\mathbb{Z}[X_{1},\ldots,X_{n},Y]_{\leq d},

and the bit-sizes of the coefficients of the polynomials in 𝒫,𝒬\mathcal{P},\mathcal{Q} are bounded by τ\tau, there exists

(2.3) cmin{2τdO(n2),2τdO(nlogd)}c\leq\min\{2^{\tau d^{O(n^{2})}},2^{\tau d^{O(n\log d)}}\}

such that the inequality (2.1) holds with N=(8d)2(n+7)N=(8d)^{2(n+7)}.

Remark 1 (Sharpness).

The inequality (2.2) is nearly tight. The following slight modification of examples given in  [45, Page 2] or [20, Example 15] shows that right hand side of inequality (2.2) cannot be made smaller than dnd^{n}. The constants (88 in the base and 1414 in the exponent) in our bound can possibly be improved (for example by using a better estimation in the inequality (5.1) in Proposition 5.1 and using a slightly more accurate degree bound). However, this would lead to a much more unwieldy statement which we prefer to avoid. The coefficient 22 of nn in the exponent however seems inherent to our method.

Example 1.

Let A:={xnx12++xn21}A:=\{x\in\mathbb{R}^{n}\mid x_{1}^{2}+\cdots+x_{n}^{2}\leq 1\} be the compact semi-algebraic set and consider the semi-algebraic functions f,g:Af,g:A\to\mathbb{R} defined by

f\displaystyle f :=|X2X1d1|++|XnXn1dn1|+|Xndn|,\displaystyle:=|X_{2}-X_{1}^{d_{1}}|+\cdots+|X_{n}-X_{n-1}^{d_{n-1}}|+|X_{n}^{d_{n}}|,
g\displaystyle g :=X12++Xn2.\displaystyle:=\sqrt{X_{1}^{2}+\cdots+X_{n}^{2}}.

It is easy to see that f1(0)=g1(0)={0}f^{-1}(0)=g^{-1}(0)=\{0\}. Then for sufficiently small |t||t| the vector x(t):=(t,td1,,td1dn1)x(t):=(t,t^{d_{1}},\ldots,t^{d_{1}\cdots d_{n-1}}) belongs to AA and we have

f(x(t))=|t|d1dn,g(x(t))=t2++t2d1dn1,\displaystyle f(x(t))=|t|^{d_{1}\cdots d_{n}},\qquad g(x(t))=\sqrt{t^{2}+\cdots+t^{2d_{1}\cdots d_{n-1}}},

which implies |g(x(t))|d1dnc|f(x(t))||g(x(t))|^{d_{1}\cdots d_{n}}\leq c\cdot|f(x(t))| for some positive constant cc. Letting d1==dn=dd_{1}=\cdots=d_{n}=d, then it follows that (f,gA)dn\mathcal{L}(f,g\mid A)\geq d^{n}.

Remark 2 (Independence from the combinatorial parameter).

An important feature of the bound in Theorem 1 is that the right hand side of inequality  (2.2) depends only on the maximum degree of the polynomials in 𝒫𝒬\mathcal{P}\cup\mathcal{Q} and is independent of the cardinalities of the sets 𝒫,𝒬\mathcal{P},\mathcal{Q}. This is not the case for the previous best known general bound due to Solernó [45] which depended on the sum of the degrees of the polynomials appearing in the descriptions of AA, ff and gg and thus implicitly on the number of polynomials involved in these descriptions. This fact, that our bound is independent of the number of polynomials, plays an important role in the applications that we discuss later in the paper. For example, it is exploited crucially in the proof of Theorem 2 (see below). Also note that the feature of being independent of combinatorial parameters is also present in some prior work that we discuss in detail in Section 3.1. But these results (notably that of Kollár [21] and also [42]) come with certain important restrictions and/or with a worse bound. In contrast, our result is completely general and nearly optimal.

Remark 3 (Separation of combinatorial and algebraic parts).

Separating the roles of combinatorial and algebraic parameters has a long history in quantitative real algebraic geometry. We include (see Section 4 below) a discussion and several prior examples of such results. The Łojasiewicz inequality is clearly a foundational result in real algebraic geometry. Hence, asking for a similar distinction in quantitative bounds on the Łojasiewicz exponent is a very natural question. Finally, the underlying idea behind making this distinction allows us to formulate and prove a version of the Łojasiewicz inequality valid over polynomially bounded o-minimal structures (see Theorem 3) which is stronger than the one known before.

Remark 4.

It is not possible to obtain a uniform bound (i.e. a bound only in terms of dd, nn and possibly the combinatorial parameters) on the constant cc in Theorem 1.

As mentioned earlier Theorem 1 leads to improvements in several applications where Łojasiewicz inequality plays an important role. We discuss some of these applications in depth in Section 6, but mention an important one right away.

2.2. Application to error bounds

Study of error bounds (defined next) is a very important topic in optimization theory and computational optimization (see for example  [43] and the references cited therein).

Definition 2.2 (Error bounds and residual function).

Let M,ERnM,E\subset\mathrm{R}^{n}. An error bound on EE with respect to MM is an inequality

(2.4) dist(x,M)ρκψ(x),xE,\displaystyle\operatorname{dist}(x,M)^{\rho}\leq\kappa\cdot\psi(x),\qquad\forall x\in E,

where ρ,κ>0\rho,\kappa>0, and ψ:MER0\psi:M\cup E\to\mathrm{R}_{\geq 0} is some function (called a residual function) such that ψ(x)=0\psi(x)=0 iff xMx\in M.

The study of error bounds was motivated by the implementation of iterative numerical optimization algorithms and the proximity of solutions to the feasible or optimal set. Thus, from the optimization point of view, the set MM in (2.4) can be the feasible set (polyhedron, a slice of the positive semi-definite cone, a basic semi-algebraic set, etc.) or the optimal set of an optimization problem, see (6.4), EE is a set of interest (e.g., iterates of an iterative algorithm or central solutions [5]), and a residual function ψ\psi measures the amount of violation of the inequalities defining MM at a given solution of EE. See [43] for other applications of error bounds in optimization.

If MM and ψ\psi are semi-algebraic and EE is a closed and bounded semi-algebraic set, then (2.4) is a special case of (1.1), and by Theorem 1, the error bound (2.4) exists with an integer ρ1\rho\geq 1.

We prove the following quantitative result.

Theorem 2.

Let MM be a basic closed semi-algebraic set defined by

(2.5) M:={xRngi(x)0,hj(x)=0,i=1,,r,j=1,,s},\displaystyle M:=\{x\in\mathrm{R}^{n}\mid g_{i}(x)\leq 0,\ h_{j}(x)=0,\ i=1,\ldots,r,\ j=1,\ldots,s\},

where gi,hjR[X1,,Xn]dg_{i},h_{j}\in\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d}, and let EE be a closed and bounded 𝒫\mathcal{P}-semi-algebraic subset of Rn\mathrm{R}^{n} with 𝒫R[X1,,Xn]d\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d}. Let ψ:MER0\psi:M\cup E\rightarrow\mathrm{R}_{\geq 0} be the semi-algebraic function defined by

(2.6) ψ(x)=j=1s|hj(x)|+i=1rmax{gi(x),0}.\displaystyle\psi(x)=\sum_{j=1}^{s}|h_{j}(x)|+\sum_{i=1}^{r}\max\{g_{i}(x),0\}.

Then there exist a positive constant κ\kappa and an integer ρ1\rho\geq 1 such that (2.4) holds, with ρ=dO(n2)\rho=d^{O(n^{2})}.

Moreover, if dimM=0\dim M=0 (i.e. MM is a finite subset of Rn\mathrm{R}^{n}), then (2.4) holds with ρ(8d)2(n+7)=dO(n)\rho\leq(8d)^{2(n+7)}=d^{O(n)}.

Remark 5.

Example 1 indicates that the upper bound on ρ\rho cannot be better than dnd^{n}.

Remark 6.

The error bounds of Theorem 2 have been stated, for the purpose of applications to optimization, only in reference to the basic semi-algebraic set (2.5) and the residual function (2.6). However, the results of Theorem 2 are still valid if we replace  (2.5) by any 𝒬\mathcal{Q}-semi-algebraic set MM and (2.6) by any 𝒬\mathcal{Q}^{\prime}-semi-algebraic residual function, where 𝒬R[X1,,Xn]d\mathcal{Q}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d} and 𝒬R[X1,,Xn,Y]dO(n)\mathcal{Q}^{\prime}\subset\mathrm{R}[X_{1},\ldots,X_{n},Y]_{\leq d^{O(n)}}, see Corollary 1.

Theorem 1 significantly improves the bound Dnc1D^{n^{c_{1}}} in [45, Theorem 7] in which DD is the sum of degrees of polynomials and c1c_{1} is universal positive integer. Furthermore, the upper bound on ρ\rho in Theorem 2 is independent of rr and ss (the number of polynomial equations and inequalities in (2.5)), which is particularly important for optimization purposes. Thus, if r,s=ω(n2)r,s=\omega(n^{2}), the first part of Theorem 2 improves the best current upper bound [30, Corollary 3.8], where

(2.7) ρmin{(d+1)(3d)n+r+s1,d(6d3)n+r1}.\displaystyle\rho\leq\min\big{\{}(d+1)(3d)^{n+r+s-1},d(6d-3)^{n+r-1}\big{\}}.

Furthermore, when MM is a finite subset of Rn\mathrm{R}^{n} and r=ω(n)r=\omega(n), the second part of Theorem 2 improves the upper bound [30, Theorem 4.1], where

(2.8) ρ(2d1)n+r+12.\displaystyle\rho\leq\frac{(2d-1)^{n+r}+1}{2}.

The improvements mentioned above are particularly relevant to non-linear semi-definite systems, non-linear semi-definite optimization, and semi-definite complementarity problems, see e.g., [28, 40], where the application of (2.7) and (2.8) would result in a doubly exponential bound since the number of inequalities needed to define the cone of positive semi-definite matrices in the vector space 𝕊n\mathbb{S}^{n} of symmetric n×nn\times n matrices with entries in R\mathrm{R} is exponential in nn. The problem of estimation of the exponent ρ\rho in the error bounds of positive semi-definite systems failing the Slater condition [15, Page 23] is posed in [28, Page 106] where it is stated “Presently, we have no idea of what this exponent ought to be except in trivial cases”. Corollary 1 quantifies the error bound exponent in [28, Proposition 6] and gives an answer to this question in the special case of polynomial mappings.

Corollary 1.

Let 𝕊+p\mathbb{S}_{+}^{p} be the cone of symmetric p×pp\times p positive semidefinite matrices (with entries in R\mathrm{R}), let MM be defined as

M:={X𝕊pgi(X)0,i=1,,r},\displaystyle M:=\{X\in\mathbb{S}^{p}\mid\ g_{i}(X)\leq 0,\ i=1,\ldots,r\},

where gi:𝕊pRg_{i}:\mathbb{S}^{p}\to\mathrm{R} is a polynomial function of degree dd, and let EE be a closed and bounded 𝒫\mathcal{P}-semi-algebraic subset of Rp2\mathrm{R}^{p^{2}} with 𝒫R[X1,,Xp2]d\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{p^{2}}]_{\leq d}. Then there exist κR\kappa\in\mathrm{R} and ρ=max{d,p}O(p4)\rho=\max\{d,p\}^{O(p^{4})} such that

dist(x,M𝕊+p)ρκmax{dist(x,M),max{λmin(x),0}}, for all xE.\displaystyle\operatorname{dist}(x,M\cap\mathbb{S}_{+}^{p})^{\rho}\leq\kappa\cdot\max\Big{\{}\operatorname{dist}(x,M),\max\{-\lambda_{\min}(x),0\}\Big{\}},\quad\mbox{ for all }x\in E.

2.3. O-minimal case

Many finiteness results of semi-algebraic geometry generalize to arbitrary o-minimal expansions of \mathbb{R} (we refer the reader to [47] and [13] for the definition of o-minimal structures and the corresponding finiteness results). Łojasiewicz inequality does not extend to arbitrary o-minimal expansions of \mathbb{R} (for example, o-minimal expansions in which the exponential function is definable). However, Miller proved in [38, Theorem 5.4, page 94] that in a polynomially bounded o-minimal expansion of \mathbb{R}, Łojasiewicz inequality holds for every compact definable set AA and definable functions f,g:Af,g:A\rightarrow\mathbb{R} with f1(0)g1(0)f^{-1}(0)\subset g^{-1}(0) (using the same notation as in Theorem 1 above). (An o-minimal expansion of \mathbb{R} is polynomially bounded if for every definable function f:f:\mathbb{R}\rightarrow\mathbb{R}, there exist N,cN\in\mathbb{N},c\in\mathbb{R}, such that |f(x)|<xN|f(x)|<x^{N} for all x>cx>c.) However, it is not possible to give a meaningful quantitative version of Miller’s result in such a general context.

2.3.1. Extension of the notion of combinatorial complexity to arbitrary o-minimal structures

Although the notion of algebraic complexity in the context of general o-minimal structure does not make sense in general – one can still talk of combinatorial complexity [3]. The following result is illustrative (see also Proposition 5.4 for another example) and can be obtained by combining [3, Theorem 2.3] and the approximation theorem proved in [16].

Fix an o-minimal expansion of the \mathbb{R} and suppose that 𝒜\mathcal{A} is a definable family of closed subsets of n\mathbb{R}^{n}. Then there exists a constant C=C(𝒜)>0C=C(\mathcal{A})>0 having the following property. Suppose that 𝒮𝒜\mathcal{S}\subset\mathcal{A} is a finite subset and SS a subset of n\mathbb{R}^{n} belonging to the Boolean algebra of subsets of n\mathbb{R}^{n} generated by 𝒮\mathcal{S} such that

(2.9) ibi(S)Csn,\sum_{i}b_{i}(S)\leq C\cdot s^{n},

where s=card(𝒮)s=\mathrm{card}(\mathcal{S}) and bi()b_{i}(\cdot) denotes the ii-th Betti number [7]. Notice that this bound does depend on the combinatorial parameter ss. Note also that it follows from Hardt’s triviality theorem for o-minimal structures [13] that the Betti numbers of the sets appearing in any definable family are bounded by a constant (depending on the family). However, the family of sets SS to which the inequality (2.9) applies is not necessarily a definable family.

In view of the inequality (2.9) it is an interesting question whether one can prove a quantitative version of Miller’s result with a uniform bound on the Łojasiewicz exponent in the same setting as above – so that the bound applies to a family (not necessarily definable) of definable sets SS and functions f,gf,g simultaneously – as in inequality (2.9).

2.3.2. Łojasiewicz inequality in polynomially bounded o-minimal structures

For the rest of this section we fix a polynomially bounded o-minimal expansion of \mathbb{R}. Before stating our theorem we need to use a notation and a definition.

Notation 1.

A definable family of subsets of n\mathbb{R}^{n} parametrized by a definable set AA, is a definable subset 𝒜A×n\mathcal{A}\subset A\times\mathbb{R}^{n}. For aAa\in A, we will denote by 𝒜a=π2(π11(a)𝒜)\mathcal{A}_{a}=\pi_{2}(\pi_{1}^{-1}(a)\cap\mathcal{A}), where π1:A×nA\pi_{1}:A\times\mathbb{R}^{n}\rightarrow A, π2:A×nn\pi_{2}:A\times\mathbb{R}^{n}\rightarrow\mathbb{R}^{n} are the two projection maps. We will often abuse notation and refer to the definable family by 𝒜\mathcal{A}.

Definition 2.3.

Given a definable family 𝒜\mathcal{A} of subsets of n\mathbb{R}^{n} parametrized by a definable set AA, and a finite subset AAA^{\prime}\subset A, we call a subset SnS\subset\mathbb{R}^{n} to be a (𝒜,A)(\mathcal{A},A^{\prime})-set if it belongs to the Boolean algebra of subsets of n\mathbb{R}^{n} generated by the tuple (𝒜a)aA(\mathcal{A}_{a})_{a\in A^{\prime}}. We will call a subset SnS\subset\mathbb{R}^{n}, a 𝒜\mathcal{A}-set if SS is a (𝒜,A)(\mathcal{A},A^{\prime})-set for some finite set AAA^{\prime}\subset A. If the graph of a definable function ff is a 𝒜\mathcal{A}-set we will call ff a 𝒜\mathcal{A}-function. (Note that the family of 𝒜\mathcal{A}-sets is in general not a definable family of subsets of n\mathbb{R}^{n}.)

Example 2.

If we take the o-minimal structure sa\mathbb{R}_{\mathrm{sa}} of semi-algebraic sets, then for each fixed dd, the family of semi-algebraic sets which are 𝒫\mathcal{P}-semi-algebraic sets where 𝒫\mathcal{P} varies over all finite subsets of [X1,,Xn]d\mathbb{R}[X_{1},\ldots,X_{n}]_{\leq d} is an example of a family of 𝒜\mathcal{A}-sets for an appropriately chosen 𝒜\mathcal{A}. Note that this family is not a semi-algebraic family.

Example 2 suggests a way to obtain a quantitative Łojasiewicz inequality valid over any polynomially bounded o-minimal structure.

Theorem 3 (Łojasiewicz inequality for 𝒜\mathcal{A}-sets and \mathcal{B}-functions for any pair of definable families 𝒜\mathcal{A} and \mathcal{B}).

Let 𝒜\mathcal{A} be a definable family of subsets of n\mathbb{R}^{n} parametrized by the definable set AA, and let \mathcal{B} be a definable family of subsets of n+1\mathbb{R}^{n+1} parametrized by the definable set BB.

Then there exists N=N(𝒜,)>0N=N(\mathcal{A},\mathcal{B})>0 having the following property. For any triple of finite sets (A,B,B′′)(A^{\prime},B^{\prime},B^{\prime\prime}) with AA,B,B′′BA^{\prime}\subset A,B^{\prime},B^{\prime\prime}\subset B, there exists c=c(A,B,B′′)c=c(A,B^{\prime},B^{\prime\prime})\in\mathbb{R} such that for each closed and bounded (𝒜,A)(\mathcal{A},A^{\prime})-set SS, a (,B)(\mathcal{B},B^{\prime})-set FF, and a (,B′′)(\mathcal{B},B^{\prime\prime})-set GG, such that F,GF,G are graphs of definable functions f,g:nf,g:\mathbb{R}^{n}\rightarrow\mathbb{R} continuous on SS with f|S1(0)g|S1(0)f|_{S}^{-1}(0)\subset g|_{S}^{-1}(0), and for all xSx\in S,

|g(x)|Nc|f(x)|.|g(x)|^{N}\leq c\cdot|f(x)|.
Remark 7 (Theorem 3 generalizes Theorem 5.4 (page 94) in [38]).

Notice that as in Theorem 1, the combinatorial parameter (namely, card(ABB′′)\mathrm{card}(A^{\prime}\cup B^{\prime}\cup B^{\prime\prime})) plays no role. It is also more general than the Łojasiewicz inequality in [38] as the inequality holds with the same value of NN for a large, potentially infinite family (not necessarily definable) of triples (S,f,g)(S,f,g) and not just for one triple as in the result in [38].

2.4. Outline of the proofs of the main theorems

We now outline the key ideas behind the proofs of the theorems stated in the previous section.

Our proof of Theorem 1 follows closely the proof of the similar qualitative statement in [10] with certain important modifications. One main tool that we use to obtain our quantitative bound is a careful analysis of the degrees of certain polynomials appearing in the output of an algorithm (Algorithm 14.6 (Block Elimination) described in the book [7]. 111We refer to the posted online version of the book because it contains certain degree estimates which are more precise than in the printed version. This algorithm is an intermediate algorithm for the effective quantifier elimination algorithm described in [7], and takes as input a finite set of polynomials 𝒫R[Y,X]\mathcal{P}\subset\mathrm{R}[Y,X], and produces as output a finite set BElimX(𝒫)R[Y]\mathrm{BElim}_{X}(\mathcal{P})\subset\mathrm{R}[Y] having the property that for each connected component CC of the realization of each realizable sign condition (see Notation 2) the set of sign conditions realized by 𝒫(y,X))\mathcal{P}(y,X)) is constant as yy varies over CC. The precise mathematical statement describing the above property of the output of the Block Elimination algorithm (including a bound on the degrees of the polynomials output) is summarized in Proposition 5.1. The proof of the bound on degrees borrows heavily from the complexity analysis of the algorithm that already appears in [7] but with an added part corresponding to the last step of the algorithm. This last step uses another algorithm (namely, Algorithm 11.54 (Restricted Elimination) in [7]) and we use the complexity analysis of this algorithm as well.

We also need a quantitative statement on the growth of a semi-algebraic function of one variable at infinity whose graph is defined by polynomials of a given degree. It is important for us that the growth is bounded only by the upper bound on the degree and not on the size of the formula describing the graph. This is proved in Lemma 5.1.

Note that the technique of utilizing complexity estimates of algorithms to prove quantitative bounds in real algebraic geometry is not altogether new. For example, similar ideas have been used to prove a quantitative curve selection lemma [9], and bounds on the radius of a ball guaranteed to intersect every connected component of a given semi-algebraic set [8], amongst other such results.

Theorem 2 is an application of Theorem 1 with one key extra ingredient. We prove (see Lemma 5.3) that if MM is a 𝒫\mathcal{P}-semi-algebraic set with 𝒫R[X1,,Xn]d\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d}, then the graph of the distance function, dist(,M){\rm dist}(\cdot,M), can be described by a quantifier-free formula involving polynomials having degrees at most dO(n)d^{O(n)}. A more naive approach (for example that in [45]) involving quantifier elimination would involve elimination of two blocks of quantified variables with a quantifier alternation and would lead to a degree bound of dO(n2)d^{O(n^{2})}. The formula describing the graph of dist(,M){\rm dist}(\cdot,M) that we obtain is very large from the point of view of combinatorial complexity (compared to the more naive approach), but with a better degree bound. We leverage now the fact that our bound on the Łojasiewicz exponent is independent of the combinatorial parameter, and apply Theorem 1 to obtain the stated result.

The proof of Theorem 3 is similar to that of proof Theorem 1 with the following important difference. Proposition 5.1 which plays a key role in the proof of Theorem 1 is replaced by a quantitative version of the existence theorem for cylindrical definable decomposition adapted to finite sub-families of a family \mathcal{F} of definable subsets of Rn\mathrm{R}^{n} in any o-minimal structure. The important quantitative property that we need is not the size of the decomposition but the fact that each cell of the decomposition is determined in a certain fixed definable way from a certain finite number, N(n)N(n), of the sets of the given finite sub-family of \mathcal{F} (the key point being that the number N(n)N(n) is independent of the cardinality of the finite sub-family). The existence of such decompositions in o-minimal structures was first observed in [3, Theorem 2.5] (see Proposition 5.5 below) and it is closely related (in fact equivalent) to the fact that o-minimal structures are distal in the sense of model theory (see [46]).

The rest of the paper is organized as follows. In Section 3 we survey prior work on proving bounds on the Łojasiewicz exponent at various levels of generality, and also survey prior work on proving error bounds. In Section 4, in order to put the current paper in context, we include a discussion of the role that the separation of combinatorial and algebraic complexity has played in quantitative real algebraic geometry. In Section 5 we prove the main theorems after introducing the necessary definitions and preliminary results. In Section 6, we discuss some further applications of our main theorem. Finally, in Section 7 we end with some open problems.

3. Prior and related work

3.1. Prior results on Łojasiewicz inequality

Solernó [45, Theorem 3] proved that (1.1) holds with ρ=Dc1n\rho=D^{c_{1}n}, in which c1c_{1} is a universal constant and DD is an upper bound on the sum of the degrees of polynomials in 𝒫\mathcal{P} and 𝒬\mathcal{Q}. Since DD is an upper bound on the sum of the degrees of polynomials in 𝒫\mathcal{P} and 𝒬\mathcal{Q}, the bound in [45, Theorem 3] depends implicitly on the cardinalities of 𝒫\mathcal{P} and 𝒬\mathcal{Q} (unlike Theorem 1). In the case of polynomials with integer coefficients, Solernó [45, Theorem 3 (ii)] also proves an upper bound of 2τDc2n22^{\tau{D^{c_{2}\cdot n^{2}}}} on the constant cc (following the same notation as in (1.1)) where DD is an upper bound on the sum of the degrees of polynomials in 𝒫\mathcal{P} and 𝒬\mathcal{Q} and c2c_{2} is a universal constant. This bound should be compared with inequality (2.3) in Theorem 1.

In a series of papers [22, 24, 25] Kurdyka, Spodzieja and Szlachcińska proved several quantitative results on Łojasiewicz inequality. We summarize them as follows. Let SnS\subset\mathbb{R}^{n} be a closed semi-algebraic set, and let

S=S1Sk\displaystyle S=S_{1}\cup\cdots\cup S_{k}

be a decomposition [10] of SS into kk closed basic 𝒫i\mathcal{P}_{i}-semi-algebraic subsets SiS_{i} with 𝒫i[X1,,Xn]di\mathcal{P}_{i}\subset\mathbb{R}[X_{1},\ldots,X_{n}]_{\leq d_{i}}, each involving rir_{i} polynomial inequalities. Let r(S)r(S) be the minimum of max{r1,,rk}\max\{r_{1},\ldots,r_{k}\} over all possible decompositions of SS and let deg(S)\deg(S) denote the minimum of max{d1,,dk}\max\{d_{1},\ldots,d_{k}\} over all decompositions for which rir(S)r_{i}\leq r(S). Further, let F:SsF:S\to\mathbb{R}^{s} be a continuous semi-algebraic mapping and suppose that 𝟎S\mathbf{0}\in S and F(𝟎)=𝟎F(\mathbf{0})=\mathbf{0}. Then for every ε>0\varepsilon>0 there exist [24, Corollary 2.2] (see also [25]) c>0c>0 and an integer

(3.1) 1ρd¯(6d¯3)n+s+r¯1\displaystyle 1\leq\rho\leq\bar{d}(6\bar{d}-3)^{n+s+\bar{r}-1}

such that

(3.2) F(x)cdist(x,F1(𝟎)S)ρ,xS,x<ε,\displaystyle\|F(x)\|\geq c\cdot\operatorname{dist}\big{(}x,F^{-1}(\mathbf{0})\cap S\big{)}^{\rho},\quad\forall x\in S,\ \|x\|<\varepsilon,

where

d¯\displaystyle\bar{d} =max{deg(S),deg(graph(F))},\displaystyle=\max\big{\{}\deg(S),\deg(\mathrm{graph}(F))\big{\}},
r¯\displaystyle\bar{r} =r(S)+r(graph(F)).\displaystyle=r(S)+r(\mathrm{graph}(F)).

If x=𝟎x=\mathbf{0} is an isolated zero of FF, then ρ((2d¯1)n+s+r¯+1)/2\rho\leq((2\bar{d}-1)^{n+s+\bar{r}}+1)/2.

Note that the above bounds do depend on the number of polynomials. Also, notice that d¯\bar{d} and r¯\bar{r} in the upper bound (3.1) are both different from dd and the number of inequalities in the semi-algebraic description of ff, gg, and AA in Theorem 1. In fact, r¯\bar{r} and d¯\bar{d} is the number of inequalities and the maximum degree of polynomials in the minimal semi-algebraic description of graph(F)\mathrm{graph}(F) and SS. It was proved in [12] that r¯\bar{r} is bounded by n(n+1)/2n(n+1)/2, but it is not clear how dd blows up for a minimal decomposition. Because of this we cannot directly compare the bound in (3.1) to that of Theorem 1 proved in the current paper.

Let f:Sf:S\to\mathbb{R} be a Nash function [10, Definition 2.9.3], where SS is a compact semi-algebraic subset of n\mathbb{R}^{n}. Osińska-Ulrych et al. [41] showed that

|f(x)|cdist(x,f1(0))2(2d¯1)3n+1,xS,\displaystyle|f(x)|\geq c\cdot\operatorname{dist}\big{(}x,f^{-1}(0)\big{)}^{2(2\bar{d}-1)^{3n+1}},\quad\forall x\in S,

in which d¯=degS(f):=max{dega(f)aS}\bar{d}=\deg_{S}(f):=\max\{\deg_{a}(f)\mid a\in S\}, and dega(f)\deg_{a}(f) is the degree of the unique irreducible P[X1,,Xn,Y]P\in\mathbb{R}[X_{1},\ldots,X_{n},Y] such that P(x,f(x))=0P(x,f(x))=0 for all xx in a connected neighborhood of aa.

Kollár [21] considered the problem of improving Solernó’s results [45]. He obtained significant improvements but under certain restrictions. More precisely, given a semi-algebraic set MM as in (2.5), with maxi{fi(x)}>0\max_{i}\{f_{i}(x)\}>0 for all xMx\in M for 0<x10<\|x\|\ll 1, and maxi{fi(𝟎)}=0\max_{i}\{f_{i}(\mathbf{0})\}=0, he proved that ([21, Theorem 4]) there exist constants c,ε>0c,\varepsilon>0 such that

(3.3) maxi{fi(x)}cxB(n1)dnfor allxMwithx<ε,\displaystyle\max_{i}\{f_{i}(x)\}\geq c\cdot\|x\|^{B(n-1)d^{n}}\ \ \text{for all}\ x\in M\ \text{with}\ \|x\|<\varepsilon,

where B(n):=(n(n/2))B(n):=\binom{n}{\lfloor(n/2)\rfloor}. Notice that the bound B(n1)dn(2d)nB(n-1)d^{n}\leq(2d)^{n} on the exponent in (3.3) is a little better than the bound in Theorem 1. It is also the case that similar to our result, Kollár’s bound is independent of the combinatorial parameters (i.e. number of polynomials occurring in the definition of MM and the number of fif_{i}’s). However, unlike Theorem 1, the pair of functions maxi{fi(x)},||||\max_{i}\{f_{i}(x)\},||\cdot|| in Kollár’s theorem is quite restrictive and so inequality (3.3) is difficult to apply directly – for instance, in applications to error bounds considered in this paper (Theorem 2). Moreover, the restriction that 𝟎\mathbf{0} has to be an isolated zero of maxi{fi(x)}\max_{i}\{f_{i}(x)\} may not be satisfied in many applications, restricting the utility of (3.3).

More recently, Osińska-Ulrych et al. [42] proved that (f,g𝔹n)d¯4n+1\mathcal{L}(f,g\mid\mathbb{B}^{n})\leq\bar{d}^{4n+1} in which d¯\bar{d} is the degree of polynomials describing ff and gg, 𝔹n\mathbb{B}^{n} is the unit ball in n\mathbb{R}^{n}, and d¯\bar{d} is a bound on the degrees of polynomials defining the graphs of f,gf,g as well as on certain polynomials giving a suitable semi-algebraic decomposition of 𝔹n\mathbb{B}^{n} adapted to f,gf,g. Note that d¯\bar{d} could be larger than the degrees of the polynomials defining f,gf,g. This bound is also independent of the combinatorial parameters but asymptotically weaker than the one in Theorem 1, and the setting is more restrictive (since the bound is not directly in terms of the degrees of the polynomials appearing in the definition of f,gf,g).

3.2. Other forms of Łojasiewicz inequality

Several other forms of Łojasiewicz inequality have appeared in the literature. Let f:Uf:U\to\mathbb{R} be a real analytic function, where UnU\subset\mathbb{R}^{n} is neighborhood of 𝟎n\mathbf{0}\in\mathbb{R}^{n}. If f(𝟎)=0f(\mathbf{0})=0 and f(𝟎)=𝟎\nabla f(\mathbf{0})=\mathbf{0}, then there exist a neighborhood UU^{\prime} of 0, a rational number ϱ<1\varrho<1, and c>0c>0 such that

(3.4) |f(x)|c|f(x)|ϱ,xU,|\nabla f(x)|\geq c\cdot|f(x)|^{\varrho},\quad\forall x\in U^{\prime},

which is known as Łojasiewicz gradient inequality. The infimum of ϱ\varrho satisfying (3.4) is called the Łojasiewicz exponent of ff, and it is denoted by ϱ(f)\varrho(f). If f[X1,,Xn]df\in\mathbb{R}[X_{1},\ldots,X_{n}]_{\leq d} and has a isolated singularity at zero, then there exists an upper bound [17]

ϱ(f)11(d1)n+1.\displaystyle\varrho(f)\leq 1-\frac{1}{(d-1)^{n}+1}.

Under a weaker condition of having non-isolated singularity at the origin, D’Acunto and Kurdyka showed [14] that

(3.5) ϱ(f)11max{d(3d4)n1,2d(3d3)n2}.\displaystyle\varrho(f)\leq 1-\frac{1}{\max\{d(3d-4)^{n-1},2d(3d-3)^{n-2}\}}.

A more general result is given by [41] for a Nash function f:Uf:U\to\mathbb{R}, where UU is a connected neighborhood of 𝟎n\mathbf{0}\in\mathbb{R}^{n}. If f(𝟎)=0f(\mathbf{0})=0 and f(𝟎)=𝟎\nabla f(\mathbf{0})=\mathbf{0}, then (3.4) holds with

ϱ(f)112(2d1)3n+1,\displaystyle\varrho(f)\leq 1-\frac{1}{2(2d-1)^{3n+1}},

where dd is the degree of the unique irreducible P[X1,,Xn,Y]P\in\mathbb{R}[X_{1},\ldots,X_{n},Y] such that P(x,f(x))=0P(x,f(x))=0 for all xUx\in U. If, in addition to the latter condition, Py(x,f(x))0\frac{\partial P}{\partial y}(x,f(x))\neq 0 for all xUx\in U, then there exists a stronger upper bound

ϱ(f)11max{2d(2d1),d(3d2)n}+1.\displaystyle\varrho(f)\leq 1-\frac{1}{\max\{2d(2d-1),d(3d-2)^{n}\}+1}.

3.3. Prior work on error bounds

Error bounds were generalized to analytic systems and basic semi-algebraic sets in [37, Theorem 2.2] and [35, Theorem 2.2] based on the analytic form of Łojasiewicz inequality and Hörmander’s results [19] but without explicit information about the exponent. Recently, an explicit upper bound with respect to MM, defined in (2.5), was given in [30] which depends exponentially on the dimension and the number of polynomial equations and inequalities, see (2.7). The upper bound (2.7) follows from the bound (3.5) and the generalized differentiation in variational analysis.

In case that MM is defined by a single convex polynomial inequality, i.e., r=1r=1, s=0s=0, then (2.4) holds with ρ(d1)n+1\rho\leq(d-1)^{n}+1 [29, Theorem 4.2]. Additionally, if there exists an xMx\in M such that g(x)<0g(x)<0, then ρ=1\rho=1 with E=nE=\mathbb{R}^{n} [29, Theorem 4.1]. More generally, there exists [11, Corollary 3.4]

(3.6) ρmin{(2d1)n+12,(n1[(n1)/2])dn}\displaystyle\rho\leq\min\Bigg{\{}\frac{(2d-1)^{n}+1}{2},\binom{n-1}{[(n-1)/2]}d^{n}\Bigg{\}}

such that the error bound (2.4) holds, where

ψ(x)=maxi{1,,r}(max{gi(x),0}).\displaystyle\psi(x)=\max_{i\in\{1,\ldots,r\}}(\max\{g_{i}(x),0\}).

A complete survey of error bounds in optimization and their applications to algorithms and sensitivity analysis can be found in [28, 43].

4. Combinatorial and algebraic complexity

A key feature of the bound in Theorem 1 is that it is independent of the cardinality of 𝒫\mathcal{P} and 𝒬\mathcal{Q} and depends only on the bound on the maximum degree of the polynomials in 𝒫𝒬\mathcal{P}\cup\mathcal{Q} and nn. In fact in many quantitative results (upper bounds on various quantities) in real algebraic geometry involving a 𝒫\mathcal{P}-semi-algebraic set a fruitful distinction can be made between the dependence of the bound on the cardinality of the set 𝒫\mathcal{P} and on the maximum degrees (or some other measure of the complexity) of the polynomials in 𝒫\mathcal{P}. The former is referred to as the combinatorial part and the latter as the algebraic part of the bound (see [4]). This distinction is important in many applications (such as in discrete and computational geometry) where the algebraic part of the bounds are treated as bounded by a fixed constant and only the combinatorial part is considered interesting.

The following examples illustrate the different nature of the dependencies on the combinatorial and the algebraic parameters in quantitative bounds appearing in real algebraic geometry and put in context the key property of Theorem 1 stated in the beginning of this subsection.

  1. 1.

    (Bound on Betti numbers.) Suppose that SRnS\subset\mathrm{R}^{n} is a 𝒫\mathcal{P}- semi-algebraic set and V=Z(Q,Rn)V={\rm Z}(Q,\mathrm{R}^{n}) a real algebraic set. Suppose that dimRV=p\dim_{\mathrm{R}}V=p, and the degrees of QQ and the polynomials in 𝒫\mathcal{P} is bounded by dd. Then,

    (4.1) ibi(SV)sp(O(d))n,\sum_{i}b_{i}(S\cap V)\leq s^{p}(O(d))^{n},

    where s=card(𝒫)s=\mathrm{card}(\mathcal{P}) and bi()b_{i}(\cdot) denotes the ii-th Betti number [7]. Notice the different dependence of the bound on ss and dd.

  2. 2.

    (Quantitative curve selection lemma.) The curve selection lemma [33, 34] (see also [39]) is a fundamental result in semi-algebraic geometry. The following quantitative version of this lemma was proved in [9].

    Theorem (Quantitative Curve Selection Lemma).

    Let 𝒫R[X1,,Xn]d\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d} be a finite set, SS a 𝒫\mathcal{P}-semi-algebraic set, and xS¯x\in\overline{S}. Then, there exist t0R,t0>0t_{0}\in\mathrm{R},t_{0}>0, a semi-algebraic path φ:[0,t0)Rn\varphi:[0,t_{0})\rightarrow\mathrm{R}^{n} with

    φ(0)=x,φ((0,t0))S,\varphi(0)=x,\ \varphi((0,t_{0}))\subset S,

    such that the degree of the Zariski closure of the image of φ\varphi is bounded by

    (O(d))4n+3.(O(d))^{4n+3}.

    Notice that the bound on the degree of the image of the curve φ\varphi in the above theorem has no combinatorial part, i.e., there is no dependence on the cardinality of 𝒫\mathcal{P} (unlike the bound in (4.1)).

  3. 3.

    (Effective quantifier elimination) Quantifier elimination is a key property of the theory of the reals, and has been studied widely from the complexity view-point. The following quantitative version appears in [6].

    Theorem 4 (Quantifier Elimination).

    Let 𝒫R[X[1],,X[ω],Y]d\mathcal{P}\subset\mathrm{R}[X_{[1]},\ldots,X_{[\omega]},Y]_{\leq d} be a finite set of ss polynomials, where X[i]X_{[i]} is a block of kik_{i} variables, and YY a block of \ell variables. Let

    Φ(Y)=(Q1X[1])(QωX[ω])Ψ(X[1],,X[ω],Y)\Phi(Y)=(Q_{1}X_{[1]})\cdots(Q_{\omega}X_{[\omega]})\Psi(X_{[1]},\ldots,X_{[\omega]},Y)

    be a quantified-formula, with Qi{,}Q_{i}\in\{\exists,\forall\} and Ψ\Psi a 𝒫\mathcal{P}-formula. Then there exists a quantifier-free formula

    Ψ(Y)=i=1Ij=1Ji(n=1Nijsign(Pijn(Y))=σijn),\Psi(Y)=\bigvee_{i=1}^{I}\bigwedge_{j=1}^{J_{i}}(\bigvee_{n=1}^{N_{ij}}\mbox{\bf sign}(P_{ijn}(Y))=\sigma_{ijn}),

    where Pijn(Y)P_{ijn}(Y) are polynomials in the variables YY, σijn{0,1,1}\sigma_{ijn}\in\{0,1,-1\},

    I\displaystyle I \displaystyle\leq s(kω+1)(k1+1)(+1)dO(kω)O(k1)O(),\displaystyle s^{(k_{\omega}+1)\cdots(k_{1}+1)(\ell+1)}d^{O(k_{\omega})\cdots O(k_{1})O(\ell)},
    Ji\displaystyle J_{i} \displaystyle\leq s(kω+1)(k1+1)dO(kω)O(k1),\displaystyle s^{(k_{\omega}+1)\cdots(k_{1}+1)}d^{O(k_{\omega})\cdots O(k_{1})},
    Nij\displaystyle N_{ij} \displaystyle\leq dO(kω)O(k1),\displaystyle d^{O(k_{\omega})\cdots O(k_{1})},

    equivalent to Φ\Phi, and the degrees of the polynomials PijkP_{ijk} are bounded by dO(kω)O(k1)d^{O(k_{\omega})\cdots O(k_{1})}.

    Moreover, if the polynomials in 𝒫\mathcal{P} have coefficients in \mathbb{Z} with bit sizes bounded by τ\tau, the polynomials PijkP_{ijk} also have integer coefficients with bit-sizes bounded by τdO(kω)O(k1)\tau d^{O(k_{\omega})\cdots O(k_{1})}.

    Notice that the bound on the degrees of the polynomials appearing in the quantifier-free formula is independent of the combinatorial parameter s=card(𝒫)s=\mathrm{card}(\mathcal{P}). This fact will play a key role in the proof of the main theorem (Theorem 1) below.

5. Proofs of the main results

5.1. Proof of Theorem 1

Before proving Theorem 1 we need some preliminary results.

5.1.1. Cylindrical definable decomposition

The notion of cylindrical definable decomposition (introduced by Łojasiewicz [33, 32]) plays a very important in semi-algebraic and more generally o-minimal geometry [13]. We include its definition below for the sake of completeness and also for fixing notation that will be needed later.

Definition 5.1 (Cylindrical definable decomposition).

Fixing the standard basis of Rn\mathrm{R}^{n}, we identify for each i,1ini,1\leq i\leq n, Ri\mathrm{R}^{i} with the span of the first ii basis vectors. Fixing an o-minimal expansion of R\mathrm{R}, a cylindrical definable decomposition of R\mathrm{R} is an 11-tuple (𝒟1)(\mathcal{D}_{1}), where 𝒟1\mathcal{D}_{1} is a finite set of subsets of R\mathrm{R}, each element being a point or an open interval, which together gives a partition of R\mathrm{R}. A cylindrical definable decomposition of Rn\mathrm{R}^{n} is an nn-tuple (𝒟1,,𝒟n)(\mathcal{D}_{1},\ldots,\mathcal{D}_{n}), where each 𝒟i\mathcal{D}_{i} is a decomposition of Ri\mathrm{R}^{i}, (𝒟1,,𝒟n1)(\mathcal{D}_{1},\ldots,\mathcal{D}_{n-1}) is a cylindrical decomposition of Rn1\mathrm{R}^{n-1}, and 𝒟n\mathcal{D}_{n} is a finite set of definable subsets of Rn\mathrm{R}^{n} (called the cells of 𝒟n\mathcal{D}_{n}) giving a partition of Rn\mathrm{R}^{n} consisting of the following: for each C𝒟n1C\in\mathcal{D}_{n-1}, there is a finite set of definable continuous functions fC,1,,fC,NC:CRf_{C,1},\ldots,f_{C,N_{C}}:C\rightarrow\mathrm{R} such that fC1<<fC,NCf_{C_{1}}<\cdots<f_{C,N_{C}}, and each element of 𝒟n\mathcal{D}_{n} is either the graph of a function fC,if_{C,i} or of the form

  1. (a)

    {(x,t)xC,t<fC,1(x)}\{(x,t)\;\mid\;x\in C,t<f_{C,1}(x)\},

  2. (b)

    {(x,t)xC,fC,i(x)<t<fC,i+1(x)}\{(x,t)\;\mid\;x\in C,f_{C,i}(x)<t<f_{C,i+1}(x)\},

  3. (c)

    {(x,t)xC,fC,NC(x)<t}\{(x,t)\;\mid\;x\in C,f_{C,N_{C}}(x)<t\},

  4. (d)

    {(x,t)xC}\{(x,t)\;\mid\;x\in C\}

(the last case arising is if the set of functions {fC,i|1iNC}\{f_{C,i}|1\leq i\leq N_{C}\} is empty).

We will say that the cylindrical definable decomposition (𝒟1,,𝒟n)(\mathcal{D}_{1},\ldots,\mathcal{D}_{n}) is adapted to a definable subset SS of Rn\mathrm{R}^{n}, if for each C𝒟nC\in\mathcal{D}_{n}, CSC\cap S is either equal to CC or empty.

In the semi-algebraic case we will refer to a cylindrical definable decomposition by cylindrical algebraic decomposition.

In the semi-algebraic case we will use the following extra notion.

5.1.2. Sign conditions

Notation 2 (Sign conditions and their realizations).

Let 𝒫\mathcal{P} be a finite subset of R[X1,,Xn]\mathrm{R}[X_{1},\ldots,X_{n}]. For σ{0,1,1}𝒫\sigma\in\{0,1,-1\}^{\mathcal{P}}, we call the formula P𝒫(sign(P)=σ(P))\bigwedge_{P\in\mathcal{P}}(\mbox{\bf sign}(P)=\sigma(P)) to be a sign condition on 𝒫\mathcal{P} and call its realization the realization of the sign condition σ\sigma. We say that a sign condition is realizable if its realization is not empty.

We denote by Cc(𝒫)\mathrm{Cc}(\mathcal{P}) the set of semi-algebraically connected components of the realizations of each realizable sign condition on 𝒫\mathcal{P}.

We say that a cylindrical algebraic decomposition 𝒟=(𝒟1,,𝒟n)\mathcal{D}=(\mathcal{D}_{1},\ldots,\mathcal{D}_{n}) of Rn\mathrm{R}^{n} is adapted to 𝒫\mathcal{P} if for each cell CC of 𝒟n\mathcal{D}_{n}, and each P𝒫P\in\mathcal{P}, sign(P(x))\mbox{\bf sign}(P(x)) is constant for xCx\in C. (This implies in particular each element of Cc(𝒫)\mathrm{Cc}(\mathcal{P}) is a union of cells of 𝒟n\mathcal{D}_{n}.)

Lemma 5.1.

Let R[X1,X2]p\mathcal{F}\subset\mathrm{R}[X_{1},X_{2}]_{\leq p} be a finite set of non-zero polynomials. Let f:RRf:\mathrm{R}\rightarrow\mathrm{R} be a semi-algebraic map, such that

graph(f)={(x,f(x))xR}=C𝒞C\mathrm{graph}(f)=\{(x,f(x))\mid x\in\mathrm{R}\}=\bigcup_{C\in\mathcal{C}}C

for some subset 𝒞Cc()\mathcal{C}\subset\mathrm{Cc}(\mathcal{F}).

Then, there exist a,cRa,c\in\mathrm{R} such that for all xax\geq a,

|f(x)|cxp.|f(x)|\leq c\cdot x^{p}.
Proof.

Consider a cylindrical algebraic decomposition 𝒟=(𝒟1,𝒟2)\mathcal{D}=(\mathcal{D}_{1},\mathcal{D}_{2}) of R2\mathrm{R}^{2} (with coordinates X1,X2X_{1},X_{2}) adapted to the set \mathcal{F}.

This implies that each CCc()C\in\mathrm{Cc}(\mathcal{F}) is a union of cells of 𝒟2\mathcal{D}_{2}. Let a0<a1<<an=aa_{0}<a_{1}<\cdots<a_{n}=a be the end-points of the intervals giving the partition of R\mathrm{R} (corresponding to the X1X_{1} coordinate) in the decomposition 𝒟1\mathcal{D}_{1}.

Since the cylindrical decomposition 𝒟\mathcal{D} is adapted to \mathcal{F}, and graph(f)=C𝒞C\mathrm{graph}(f)=\bigcup_{C\in\mathcal{C}}C for some subset 𝒞Cc()\mathcal{C}\subset\mathrm{Cc}(\mathcal{F}), dim(C)1\dim(C)\leq 1 for each C𝒞C\in\mathcal{C}, since

dim(C)dim(graph(f))=1.\dim(C)\leq\dim(\mathrm{graph}(f))=1.

Hence, there exists for each C𝒞C\in\mathcal{C} a polynomial FF\in\mathcal{F}, such that F(x)=0F(x)=0 for all xCx\in C.

Also, since graph(f)=C𝒞C\mathrm{graph}(f)=\bigcup_{C\in\mathcal{C}}C and each C𝒞C\in\mathcal{C} is a union of cells of 𝒟\mathcal{D}, there exists a continuous semi-algebraic function γ:(a,)R\gamma:(a,\infty)\rightarrow\mathrm{R}, such that graph(γ)graph(f)\mathrm{graph}(\gamma)\subset\mathrm{graph}(f), and graph(γ)\mathrm{graph}(\gamma) is a cell of 𝒟2\mathcal{D}_{2}.

Let CCc(𝒞)C\in\mathrm{Cc}(\mathcal{C}) be the unique element of 𝒞\mathcal{C} which contains graph(γ)\mathrm{graph}(\gamma), and FF\in\mathcal{F} such that F(x)=0F(x)=0 for all xCx\in C.

The lemma now follows from [10, Proposition 2.6.1] noting that

deg(F)p.\deg(F)\leq p.

Notation 3 (Realizable sign conditions).

For any finite set of polynomials 𝒫R[X1,,Xk]\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{k}] we denote by

SIGN(𝒫){0,1,1}𝒫\mathrm{SIGN}(\mathcal{P})\subset\{0,1,-1\}^{\mathcal{P}}

the set of all realizable sign conditions for 𝒫\mathcal{P} over Rk\mathrm{R}^{k}, i.e.

SIGN(𝒫)={σ{0,1,1}𝒫(σ,Rn)}.\mathrm{SIGN}(\mathcal{P})=\{\sigma\in\{0,1,-1\}^{\mathcal{P}}\;\mid\;{\mathcal{R}}(\sigma,\mathrm{R}^{n})\neq\emptyset\}.
Proposition 5.1.

Let 𝒫R[Y,X]\mathcal{P}\subset\mathrm{R}[Y,X], with Y=(Y1,,Y),X=(X1,,Xk)Y=(Y_{1},\ldots,Y_{\ell}),X=(X_{1},\ldots,X_{k}) be a finite set of polynomials. Then there exists a finite subset BElimX(𝒫)R[Y]\mathrm{BElim}_{X}(\mathcal{P})\subset\mathrm{R}[Y] such that for each CCc(BElimX(𝒫))C\in\mathrm{Cc}(\mathrm{BElim}_{X}(\mathcal{P})), SIGN(𝒫(y,X)\mathrm{SIGN}(\mathcal{P}(y,X) is fixed as yy varies over CC.

If the degrees of the polynomials in 𝒫\mathcal{P} are bounded by d2d\geq 2, then the degrees of the polynomials in BElimX(𝒫)\mathrm{BElim}_{X}(\mathcal{P}) is bounded by

(5.1) 8d2(2k(2d+2)+2)(2d+3)(2d+6)2(2d+5)2k2<(8d)2k+4.8d^{2}(2k(2d+2)+2)(2d+3)(2d+6)^{2}(2d+5)^{2k-2}<(8d)^{2k+4}.
Proof.

Let BElimX(𝒫)\mathrm{BElim}_{X}(\mathcal{P}) be the set of polynomials denoted by the same formula in the output of Algorithm 14.6 (Block Elimination) in [7]. The fact that for each CCc(BElimX(𝒫))C\in\mathrm{Cc}(\mathrm{BElim}_{X}(\mathcal{P})), SIGN(𝒫(y,X)\mathrm{SIGN}(\mathcal{P}(y,X) is fixed as yy varies over CC is a consequence of Proposition 14.10 in [7].

To obtain the upper bound on the degrees of the polynomials in BElimX(𝒫)\mathrm{BElim}_{X}(\mathcal{P}), we follow the complexity analysis of Algorithm 14.6 (Block Elimination) in [7] using the same notation as in the algorithm. The algorithm first computes a set URX(𝒫)\mathrm{UR}_{X}(\mathcal{P}) whose elements are tuples v=(f,g0,,gk)v=(f,g_{0},\ldots,g_{k}) of polynomials in T,Y,ε,δT,Y,{\varepsilon},\delta (here ε{\varepsilon} and δ\delta are infinitesimals and TT is one variable). It is proved in the complexity analysis of the algorithm that the degrees of the polynomials in TT appearing in the various tuples vURX(𝒫)v\in\mathrm{UR}_{X}(\mathcal{P}) are bounded by

D=(2d+6)(2d+5)k2,D=(2d+6)(2d+5)^{k-2},

and their degrees in YY (as well as in ε,δ{\varepsilon},\delta) are bounded by

D=(2k(2d+2)+2)(2d+3)(2d+6)(2d+5)k2.D^{\prime}=(2k(2d+2)+2)(2d+3)(2d+6)(2d+5)^{k-2}.

It follows that for each P𝒫P\in\mathcal{P}, and v=(f,g0,,gk)URX(𝒫)v=(f,g_{0},\ldots,g_{k})\in\mathrm{UR}_{X}(\mathcal{P}), the degree in TT of the polynomial

Pv=g0eP(g1g0,,gkg0),P_{v}=g_{0}^{e}P(\frac{g_{1}}{g_{0}},\ldots,\frac{g_{k}}{g_{0}}),

where ee is the least even integer greater than deg(P)d\deg(P)\leq d, is bounded by (d+1)D2dD(d+1)D\leq 2dD, and similarly the degree in YY of PvP_{v} is bounded by 2dD2dD^{\prime}. The same bounds apply to all polynomials in the set v\mathcal{F}_{v} introduced in the algorithm.

It now follows from the complexity analysis of Algorithm 11.54 (Restricted Elimination) in [7], that the degrees in YY of the polynomials in RElimT(f,v)\mathrm{RElim}_{T}(f,\mathcal{F}_{v}) are bounded by

2(2dD)(2dD)\displaystyle 2(2dD)(2dD^{\prime}) =\displaystyle= 8d2DD\displaystyle 8d^{2}DD^{\prime}
=\displaystyle= 8d2(2k(2d+2)+2)(2d+3)(2d+6)2(2d+5)2k2\displaystyle 8d^{2}(2k(2d+2)+2)(2d+3)(2d+6)^{2}(2d+5)^{2k-2}
\displaystyle\leq (8d2)(6kd)(4d)(5d)2(4d)2k2\displaystyle(8d^{2})\cdot(6kd)\cdot(4d)\cdot(5d)^{2}\cdot(4d)^{2k-2}
=\displaystyle= 86452kd6(4d)2k2\displaystyle 8\cdot 6\cdot 4\cdot 5^{2}\cdot k\cdot d^{6}\cdot(4d)^{2k-2}
=\displaystyle= 35243k(4d)2k+4\displaystyle\frac{3\cdot 5^{2}}{4^{3}}\cdot k\cdot(4d)^{2k+4}
<\displaystyle< (8d)2k+4.\displaystyle(8d)^{2k+4}.

Denoting vR[Y]\mathcal{B}_{v}\subset\mathrm{R}[Y] the set of coefficients of the various polynomials in RElimT(f,v)\mathrm{RElim}_{T}(f,\mathcal{F}_{v}) written as polynomials in ε,δ{\varepsilon},\delta, we have immediately obtain that the degrees in YY of polynomials in v\mathcal{B}_{v} are bounded by

8d2(2k(2d+2)+2)(2d+3)(2d+6)2(2d+5)2k2<(8d)2k+4.8d^{2}(2k(2d+2)+2)(2d+3)(2d+6)^{2}(2d+5)^{2k-2}<(8d)^{2k+4}.

The proposition follows since according to the algorithm

BElimX(𝒫)=vURX(𝒫)v.\mathrm{BElim}_{X}(\mathcal{P})=\bigcup_{v\in\mathrm{UR}_{X}(\mathcal{P})}\mathcal{B}_{v}.

Lemma 5.2.

Suppose that 𝒫R[Y,X]\mathcal{P}\subset\mathrm{R}[Y,X], with Y=(Y1,,Y),X=(X1,,Xk)Y=(Y_{1},\ldots,Y_{\ell}),X=(X_{1},\ldots,X_{k}) and Φ\Phi is 𝒫\mathcal{P}-formula. Then there exist subsets 𝒞,𝒞Cc(BElimX(𝒫))\mathcal{C}_{\exists},\mathcal{C}_{\forall}\subset\mathrm{Cc}(\mathrm{BElim}_{X}(\mathcal{P})), such that

((X)Φ,R)\displaystyle{\mathcal{R}}((\exists X)\Phi,\mathrm{R}^{\ell}) =\displaystyle= C𝒞C,\displaystyle\bigcup_{C\in\mathcal{C}_{\exists}}C,
(5.2) ((X)Φ,R)\displaystyle{\mathcal{R}}((\forall X)\Phi,\mathrm{R}^{\ell}) =\displaystyle= C𝒞C.\displaystyle\bigcup_{C\in\mathcal{C}_{\forall}}C.
Proof.

The lemma follows from the fact that for each CCc(BElimX(𝒫))C\in\mathrm{Cc}(\mathrm{BElim}_{X}(\mathcal{P})), the set SIGN(𝒫(y,X)\mathrm{SIGN}(\mathcal{P}(y,X) is fixed as yy varies over CC (Proposition 5.1), and the observation that for each yRy\in\mathrm{R}^{\ell}, the truth or falsity of each of the formulas (X)Φ(y,X),(X)Φ(y,X)(\exists X)\Phi(y,X),(\forall X)\Phi(y,X) is determined by the set SIGN(𝒫(y,X)\mathrm{SIGN}(\mathcal{P}(y,X). ∎

The following proposition is the key ingredient in the proof of Theorem 1. It can be viewed as a quantitative version of Proposition 2.6.4 in [10] (which is not quantitative). Our proof is similar in spirit to the proof of Proposition 2.6.4 in [10] but differs at certain important points making it possible to achieve the quantitative bound claimed in the proposition.

Proposition 5.2.

Let d2d\geq 2, 𝒫R[X1,,Xn]d\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d} and

𝒫f,𝒫gR[X1,,Xn,Y]d\mathcal{P}_{f},\mathcal{P}_{g}\subset\mathrm{R}[X_{1},\ldots,X_{n},Y]_{\leq d}

be finite sets of polynomials. Let ARnA\subset\mathrm{R}^{n} be a closed 𝒫\mathcal{P}-semi-algebraic set, f:ARf:A\rightarrow\mathrm{R} a continuous semi-algebraic function whose graph is a 𝒫f\mathcal{P}_{f}-semi-algebraic set, and g:{xAf(x)0}Rg:\{x\in A\;\mid\;f(x)\neq 0\}\rightarrow\mathrm{R} a continuous semi-algebraic function whose graph is a 𝒫g\mathcal{P}_{g}-semi-algebraic set. Then there exists N(8d)2n+10N\leq(8d)^{2n+10} such that the function fNgf^{N}g extended by 0 on {xAf(x)=0}\{x\in A\;\mid\;f(x)=0\} is semi-algebraic and continuous on AA.

Proof.

Suppose that A=(Φ,Rn)A={\mathcal{R}}(\Phi,\mathrm{R}^{n}) graph(f)=(Φf,Rn+1)\mathrm{graph}(f)={\mathcal{R}}(\Phi_{f},\mathrm{R}^{n+1}), and graph(g)=(Φg,Rn+1)\mathrm{graph}(g)={\mathcal{R}}(\Phi_{g},\mathrm{R}^{n+1}), where where Φ\Phi is a 𝒫\mathcal{P}-formula, Φf\Phi_{f} is a 𝒫f\mathcal{P}_{f}-formula, and Φg\Phi_{g} is a 𝒫g\mathcal{P}_{g}-formula.

For each xA,uRx\in A,u\in\mathrm{R}, we define

Ax,u={yAyx1,u|f(y)|=1}.A_{x,u}=\{y\in A\;\mid\;||y-x||\leq 1,u|f(y)|=1\}.

We define Θ(X,U,Y,V)\Theta(X,U,Y,V) to be the quantifier-free formula

 Φ(Y)(YX210)    ((V>0(UV1=0))((V<0)(UV+1=0)))    ϕf(Y,V). \halign{\hbox to\displaywidth{$\hfil\displaystyle#\hfil$}\cr 0.0pt{\hfil$\displaystyle\Phi(Y)\;\wedge\;(||Y-X||^{2}-1\leq 0)\cr 0.0pt{\hfil$\displaystyle\;\wedge\;\cr 0.0pt{\hfil$\displaystyle((V>0\;\wedge\;(UV-1=0))\;\vee\;((V<0)\;\wedge\;(UV+1=0)))\cr 0.0pt{\hfil$\displaystyle\;\wedge\;\cr 0.0pt{\hfil$\displaystyle\phi_{f}(Y,V).\crcr}}}}}}

Observe that for each xAx\in A and uRu\in\mathrm{R},

(Θ(x,u,,),Rn+1)=graph(f|Ax,u).{\mathcal{R}}(\Theta(x,u,\cdot,\cdot),\mathrm{R}^{n+1})=\mathrm{graph}(f|_{A_{x,u}}).

The semi-algebraic set Ax,uA_{x,u} is closed and bounded, and we define the semi-algebraic function

v(x,u)={0 if Ax,u=,sup{|g(y)|yAx,u} otherwise.v(x,u)=\begin{cases}0\mbox{ if }A_{x,u}=\emptyset,\\ \sup\{|g(y)|\;\mid\;y\in A_{x,u}\}\mbox{ otherwise}.\end{cases}

Let Λ0(X,U,W)\Lambda_{0}(X,U,W) denote the following the first-order (quantified) formula:

(Y,V,Z)(W0)(Θ(X,U,Y,V)ϕg(Y,Z))((Z0)(WZ))((Z0)(WZ))).\forall(Y,V,Z)(W\geq 0)\;\wedge\;(\Theta(X,U,Y,V)\;\wedge\;\phi_{g}(Y,Z))\;\Longrightarrow\\ ((Z\geq 0)\;\wedge\;(W\geq Z))\;\vee\;((Z\leq 0)\;\wedge\;(W\geq-Z))).

Finally, let Λ(X,U,W)\Lambda(X,U,W) denote the formula

(W)Λ0(X,U,W)(0WW).(\forall W^{\prime})\;\Lambda_{0}(X,U,W^{\prime})\;\Longrightarrow\;(0\leq W\leq W^{\prime}).

Notice that for Λ(x,u,w)\Lambda(x,u,w) is true if and only if w=v(x,u)w=v(x,u). Also notice that for each xAx\in A, Λ(x,U,W)\Lambda(x,U,W) is equivalent to a formula

(Y,V,W,Z)Ψx(Y,V,Z,U,W,W)\forall(Y,V,W^{\prime},Z)\;\Psi_{x}(Y,V,Z,U,W,W^{\prime})

to a (n+5)(n+5)-ary 𝒫x\mathcal{P}_{x}-formula Ψx\Psi_{x}, for some finite set 𝒫xR[Y,V,Z,U,W,W]max(d,2)\mathcal{P}_{x}\subset\mathrm{R}[Y,V,Z,U,W,W^{\prime}]_{\leq\max(d,2)} (since we assume d2d\geq 2).

Let

𝒬x=BElimY,V,W,Z(𝒫x)R[U,W]\mathcal{Q}_{x}=\mathrm{BElim}_{Y,V,W^{\prime},Z}(\mathcal{P}_{x})\subset\mathrm{R}[U,W]

(see Proposition 5.1).

Then, using the degree bound in Proposition 5.1 we have that for each Q𝒬xQ\in\mathcal{Q}_{x}, deg(Q)<(8d)2(n+3)+4=(8d)2n+10\deg(Q)<(8d)^{2(n+3)+4}=(8d)^{2n+10}.

It now follows from Lemma 5.1 that there exists c=c(x)c=c(x), such that for all uc(x)u\geq c(x),

|v(x,u)|cup,|v(x,u)|\leq c\cdot u^{p},

with p<(8d)2n+10p<(8d)^{2n+10}.

This means that

|f(y)|p|g(y)|c(x)|f(y)|^{p}|g(y)|\leq c(x)

on {yAf(y)0 and yx1}\{y\in A\;\mid\;f(y)\neq 0\ \mbox{ and }\ ||y-x||\leq 1\} for |f(y)||f(y)| sufficiently small. The function fNgf^{N}g extended by 0 is then semi-algebraic and continuous at xx, where N=p+1(8d)2n+10N=p+1\leq(8d)^{2n+10}. This completes the proof. ∎

Theorem 5.

Let d2d\geq 2, and

𝒫R[X1,,Xn]d,𝒫f,𝒫gR[X1,,Xn,Y]d.\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d},\mathcal{P}_{f},\mathcal{P}_{g}\subset\mathrm{R}[X_{1},\ldots,X_{n},Y]_{\leq d}.

Let ARnA\subset\mathrm{R}^{n} be a closed 𝒫\mathcal{P}-semi-algebraic set, f,g:ARf,g:A\rightarrow\mathrm{R} be continuous semi-algebraic functions whose graphs are 𝒫f\mathcal{P}_{f}-semi-algebraic respectively 𝒫g\mathcal{P}_{g}-semi-algebraic set, and such that f1(0)g1(0)f^{-1}(0)\subset g^{-1}(0). Then there exist N=(8d)2(n+7)N=(8d)^{2(n+7)} and a continuous semi-algebraic function h:ARh:A\rightarrow\mathrm{R} such that gN=hfg^{N}=hf on AA.

Proof.

Suppose that A=(Φ,Rn)A={\mathcal{R}}(\Phi,\mathrm{R}^{n}) graph(f)=(Φf,Rn+1)\mathrm{graph}(f)={\mathcal{R}}(\Phi_{f},\mathrm{R}^{n+1}), and graph(g)=(Φg,Rn+1)\mathrm{graph}(g)={\mathcal{R}}(\Phi_{g},\mathrm{R}^{n+1}), where where Φ\Phi is a 𝒫\mathcal{P}-formula, Φf\Phi_{f} is a 𝒫f\mathcal{P}_{f}-formula, and Φg\Phi_{g} is a 𝒫g\mathcal{P}_{g}-formula.

Let A~={(x,f(x),g(x))|xA}Rn+2\widetilde{A}=\{(x,f(x),g(x))|x\in A\}\subset\mathrm{R}^{n+2}. The function 1/f1/f is continuous semi-algebraic on {(x,u,v)A~g(x)0}\{(x,u,v)\in\widetilde{A}\;\mid\;g(x)\neq 0\}, and its graph is defined by the formula

Φg(X,V)(V0)Φf(X,U)(UW1=0)\Phi_{g}(X,V)\wedge(V\neq 0)\wedge\Phi_{f}(X,U)\wedge(UW-1=0)

Moreover using Proposition 5.2 there exists N(8d)2(n+2)+10=(8d)2(n+7)N\leq(8d)^{2(n+2)+10}=(8d)^{2(n+7)} such that the function h~:A~R\widetilde{h}:\widetilde{A}\rightarrow\mathrm{R} defined by

h~(x,u,v)={0 if f(x)=0,gN(x)/f(x) if f(x)0.\widetilde{h}(x,u,v)=\begin{cases}0\mbox{ if }f(x)=0,\\ g^{N}(x)/f(x)\mbox{ if }f(x)\neq 0.\end{cases}

Since h~\widetilde{h} does not depend on u,vu,v, we get a continuous and semi-algebraic function h(x)=h~(x,f(x),g(x))h(x)=\widetilde{h}(x,f(x),g(x)) on AA, and gN=hfg^{N}=hf. ∎

Proof of Theorem 1.

In order to prove inequality (2.2), use Theorem 5 with c=supxA|h(x)|c=\sup_{x\in A}|h(x)|, noticing that cc exists since AA is assumed to be closed and bounded.

We now prove inequality (2.3). The set of c,c>0c\subset\mathbb{R},c>0 for which inequality (2.1) holds for all xAx\in A is defined by

Θ(C):=(C>0)(X,U,V)(ϕ(X)ϕf(X,U)ϕg(X,V))(V2NC2U2)),\Theta(C):=(C>0)\wedge\left(\forall X,U,V)\left(\phi(X)\wedge\phi_{f}(X,U)\wedge\phi_{g}(X,V)\right)\Longrightarrow(V^{2N}\leq C^{2}\cdot U^{2})\right),

where ϕ\phi is a 𝒫\mathcal{P}-formula describing AA, and ϕf,ϕg\phi_{f},\phi_{g} are 𝒬\mathcal{Q}-formulas describing the graphs of ff and gg and N=(8d)2(n+7)N=(8d)^{2(n+7)}.

Using Theorem 4 we obtain that Θ(C)\Theta(C) is equivalent to a quantifier-free formula Θ~(C)\widetilde{\Theta}(C) such that the bit sizes of the coefficients of the polynomials appearing in Θ~(C)\widetilde{\Theta}(C) is bounded by τdO(n2)\tau d^{O(n^{2})} and their degrees are bounded by dO(n2)d^{O(n^{2})}. Now using Cauchy’s bound ([7, Lemma 10.2]), the largest real root amongst the real roots of the polynomials appearing in Θ~(C)\widetilde{\Theta}(C) is bounded by 2τdO(n2)2^{\tau d^{O(n^{2})}}. It follows that there exists c=2τdO(n2)c=2^{\tau d^{O(n^{2})}} for which the inequality (2.1) holds.

Using the repeated squaring technique (see below) at the cost of introducing O(nlogd)O(n\log d) new variables, it is possible write another universally quantified formula, namely

Θ(C):=(C>0)(T1,,TM,,X,U,V)(ϕ(X)ϕf(X,U)ϕg(X,V)(T1=V)(T2=T12)(TM=TM12))(TM2C2U2)),\Theta^{\prime}(C):=(C>0)\wedge(\forall T_{1},\ldots,T_{M},,X,U,V)(\phi(X)\wedge\phi_{f}(X,U)\wedge\phi_{g}(X,V)\wedge\cr(T_{1}=V)\wedge(T_{2}=T_{1}^{2})\wedge\cdots\wedge(T_{M}=T_{M-1}^{2}))\Longrightarrow(T_{M}^{2}\leq C^{2}\cdot U^{2})),

equivalent to Θ(C)\Theta(C) in which all the polynomials appearing have degrees bounded by dd (instead of dO(n)d^{O(n)} as in the formula Θ\Theta). The number of quantifier variables in the formula Θ\Theta^{\prime} equals M+n+2M+n+2, where M=O(logN)=O(nlogd)M=O(\log N)=O(n\log d).

Now using Theorem 4 and Cauchy’s bound as before we obtain a bound of 2τdO(nlogd)2^{\tau d^{O(n\log d)}} on cc. ∎

5.2. Proof of Theorem 2

First, we need the following lemma.

Lemma 5.3.

Let 𝒫R[X1,,Xn]d\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d} and SRnS\subset\mathrm{R}^{n} a 𝒫\mathcal{P}-semi-algebraic set. Then there exists 𝒬R[X1,,Xn,U]\mathcal{Q}\subset\mathrm{R}[X_{1},\ldots,X_{n},U] such that the graph of the function dist(,S):RnR\operatorname{dist}(\cdot,S):\mathrm{R}^{n}\rightarrow\mathrm{R} is a 𝒬\mathcal{Q}-semi-algebraic set and maxQ𝒬deg(Q)=dO(n)\max_{Q\in\mathcal{Q}}\deg(Q)=d^{O(n)}.

Before proving Lemma 5.3 we need a new notion (that of Thom encoding of real roots of a polynomial) that will be needed in the proof. The following proposition appears in [7, Proposition 2.36].

Proposition 5.3 (Thom’s Lemma).

[7, Proposition 2.36] Let PR[X]P\subset\mathrm{R}[X] be a univariate polynomial, Der(P){\rm Der}(P) the set of derivatives of PP and σ{1,0,1}Der(P)\sigma\in\{-1,0,1\}^{{\rm Der}(P)}. Then (σ,R){\mathcal{R}}(\sigma,\mathrm{R}) is either empty, a point, or an open interval.

Note that it follows immediately from Proposition 5.3, that for any PR[X]P\in\mathrm{R}[X] and xRx\in\mathrm{R} such that P(x)=0P(x)=0, the sign condition σ\sigma realized by Der(P){\rm Der}(P) at xx characterizes the root xx. We call σ\sigma the Thom encoding of the root xx of PP.

Proof of Lemma 5.3.

Let Φ\Phi be a 𝒫\mathcal{P}-formula such that (Φ,Rn)=S{\mathcal{R}}(\Phi,\mathrm{R}^{n})=S. Let

W={(x,t)yS with t=xy}.W=\{(x,t)\;\mid\;\exists y\in S\mbox{ with }t=||x-y||\}.

Then for each xRnx\in\mathrm{R}^{n},

dist(x,S)=infWx,\operatorname{dist}(x,S)=\inf W_{x},

where WxW_{x} denotes the one-dimensional fiber of WW over xx with respect to the projection along the tt coordinate.

It is also clear from the definition that WW is the image under projection along the yy coordinates) of the semi-algebraic set VRn×Rn×RV\subset\mathrm{R}^{n}\times\mathrm{R}^{n}\times\mathrm{R} defined by

V={(x,y,t)yS and t=xy}.V=\{(x,y,t)\;\mid\;y\in S\mbox{ and }t=||x-y||\}.

Let Θ(X,Y,T)\Theta(X,Y,T) be the formula

Φ(Y)(T2=XY2)(T0).\Phi(Y)\wedge(T^{2}=||X-Y||^{2})\wedge(T\geq 0).

Then, it is clear that Θ\Theta is a 𝒫\mathcal{P}^{\prime}-formula for some finite subset 𝒫R[X,Y,T]d\mathcal{P}^{\prime}\subset\mathrm{R}[X,Y,T]_{\leq d} (assuming d2d\geq 2), and moreover (Θ,R2n+1)=V{\mathcal{R}}(\Theta,\mathrm{R}^{2n+1})=V.

Now apply Theorem 4 to the quantified formula (Y)Θ(X,Y,T)(\exists Y)\Theta(X,Y,T) and obtain a quantifier-free \mathcal{F}-formula, Θ~\widetilde{\Theta} equivalent to Θ\Theta, where \mathcal{F} is some finite subset of R[X,T]\mathrm{R}[X,T] and

D=maxFdeg(F)=dO(n).D=\max_{F\in\mathcal{F}}\deg(F)=d^{O(n)}.

Notice that (Θ~,Rn+1)=W{\mathcal{R}}(\widetilde{\Theta},\mathrm{R}^{n+1})=W, and for each xRnx\in\mathrm{R}^{n},

infWx=dist(x,S)\inf W_{x}=\operatorname{dist}(x,S)

is a real root of some polynomial in \mathcal{F}.

Let ={F1,,FN}\mathcal{F}=\{F_{1},\ldots,F_{N}\}. We denote

DerT(Fi)={Fi,Fi,,Fi(D)}{\rm Der}_{T}(F_{i})=\{F_{i},F_{i}^{\prime},\ldots,F_{i}^{(D)}\}

the set of derivatives of FiF_{i} with respect to TT.

For 1iN1\leq i\leq N, and σ{1,0,1}DerT(Fi)\sigma\in\{-1,0,1\}^{{\rm Der}_{T}(F_{i})}, denote by Ψi,σ\Psi_{i,\sigma} the quantifier-free formula,

(T)(0jD(sign(Fi(j))=σ(Fi(j))))((T)Θ~(X,T)(TT)).(\forall T)\left(\bigwedge_{0\leq j\leq D}(\mbox{\bf sign}(F_{i}^{(j)})=\sigma(F_{i}^{(j)}))\right)\Longrightarrow\left((\forall T^{\prime})\widetilde{\Theta}(X,T^{\prime})\Longrightarrow(T\leq T^{\prime})\right).

Using Theorem 4 one more time, let Ψ~i,σ\widetilde{\Psi}_{i,\sigma} be a quantifier-free formula equivalent to Ψi,σ\Psi_{i,\sigma} and let the set of polynomials appearing in Ψ~i,σ\widetilde{\Psi}_{i,\sigma} be denoted by 𝒬i,σ\mathcal{Q}_{i,\sigma}.

The semi-algebraic set (Ψ~i,σ,Rn){\mathcal{R}}(\widetilde{\Psi}_{i,\sigma},\mathrm{R}^{n}) is the set consisting of points xRnx\in\mathrm{R}^{n}, such that the dist(x,S)\operatorname{dist}(x,S) equals the (at most one) real root of the polynomial Fi(x,T)F_{i}(x,T) with Thom encoding σ\sigma. Notice that the maximum degree of polynomials in 𝒬i,σ\mathcal{Q}_{i,\sigma} is bounded by DO(1)=dO(n)D^{O(1)}=d^{O(n)}.

Finally, let

Ψ=i,σ(Ψ~i,σ(0jD(sign(Fi(j))=σ(Fi(j))))),\Psi=\bigwedge_{i,\sigma}\left(\widetilde{\Psi}_{i,\sigma}\Longrightarrow\left(\bigwedge_{0\leq j\leq D}(\mbox{\bf sign}(F_{i}^{(j)})=\sigma(F_{i}^{(j)}))\right)\right),

and

𝒬=1iN(DerT(Fi)σ{1,0,1}DerT(Fi)𝒬i,σ).\mathcal{Q}=\bigcup_{1\leq i\leq N}\left({\rm Der}_{T}(F_{i})\cup\bigcup_{\sigma\in\{-1,0,1\}^{{\rm Der}_{T}(F_{i})}}\mathcal{Q}_{i,\sigma}\right).

It is clear from the above construction that, Ψ\Psi is a 𝒬\mathcal{Q}-formula, and (Ψ,Rn+1){\mathcal{R}}(\Psi,\mathrm{R}^{n+1}) is the graph of the function dist(,S)\operatorname{dist}(\cdot,S), and the degrees of the polynomials in 𝒬\mathcal{Q} are bounded by dO(n)d^{O(n)}. This proves the lemma. ∎

Remark 8.

In [10, Proposition 2.2.8] and [45, Theorem 7], the graph of the semi-algebraic function dist(x,)\operatorname{dist}(x,\mathcal{M}) is described by the following quantified formula with two blocks of quantifiers

(5.3) (Y)(Y)(¬Ψ(Y)(Ψ(Y)(||XY||2=T2)(T0)(||XY||2T2)),(\exists Y)\ (\forall Y^{\prime})\ \ (\neg\Psi(Y^{\prime})\vee(\Psi(Y)\wedge(||X-Y||^{2}=T^{2})\wedge(T\geq 0)\wedge(||X-Y^{\prime}||^{2}\geq T^{2})),

where (Y)(\exists Y) and (Y)(\forall Y^{\prime}) are two blocks of variables each of size nn with different quantifiers. The effective quantifier elimination (Theorem 4) applied to (5.3) yields a quantifier-free formula with polynomials having degrees bounded by dO(n2)d^{O(n^{2})}, where dd is an upper bound on the degrees of the polynomials in Ψ\Psi. The formula given in Lemma 5.3 describing the graph of the same function involves polynomials of degrees at most dO(n)d^{O(n)} (though it may involve many more polynomials than the one in (5.3)) which is an improvement over the bound of dO(n2)d^{O(n^{2})} obtained by the above argument. Note that for our purposes, the degrees of the polynomials appearing in the formula is the important quantity – the number of polynomials appearing is not relevant.

Proof of Theorem 2.

It is easy to see that the residual function ψ(x)\psi(x) defined in (2.6) satisfies

dist(x,M)=0ψ(x)=j=1s|hj(x)|+i=1rmax{gi(x),0}=0.\displaystyle\operatorname{dist}(x,M)=0\ \ \Longleftrightarrow\ \ \psi(x)=\sum_{j=1}^{s}|h_{j}(x)|+\sum_{i=1}^{r}\max\{g_{i}(x),0\}=0.

The graph of ψ(x)\psi(x) can be described using a quantifier-free 𝒫\mathcal{P}-formula with 𝒫R[X1,,Xn]d\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d} as follows (note that we do not care about the cardinality of 𝒫\mathcal{P}).

To see this first observe that for all xRnx\in\mathrm{R}^{n}, and σ{0,1,1}[1,s]\sigma\in\{0,1,-1\}^{[1,s]} if sign(hj(x))=σ(j),j[1,s]\mbox{\bf sign}(h_{j}(x))=\sigma(j),j\in[1,s], then

j=1s|hj(x)|=j=1sσ(j)hj(x).\sum_{j=1}^{s}|h_{j}(x)|=\sum_{j=1}^{s}\sigma(j)h_{j}(x).

Similarly, for all xRnx\in\mathrm{R}^{n}, and τ{0,1,1}[1,r]\tau\in\{0,1,-1\}^{[1,r]} if sign(gi(x))=τ(i),i[1,r]\mbox{\bf sign}(g_{i}(x))=\tau(i),i\in[1,r], then

i=1rmax{gi(x),0}=12i=1rτ(i)(1+τ(i))gi(x).\sum_{i=1}^{r}\max\{g_{i}(x),0\}=\frac{1}{2}\sum_{i=1}^{r}\tau(i)(1+\tau(i))g_{i}(x).

It is now easy to verify that the following quantifier-free formula in n+1n+1 variables:

(5.4) σ{0,1,1}[1,s]τ{0,1,1}[1,r](j=1s(sign(hj)=σ(j))i=1r(sign(gi)=τ(i)))(Tj=1sσ(j)hj12i=1rτ(i)(1+τ(i))gi=0)\bigvee_{\begin{subarray}{c}\sigma\in\{0,1,-1\}^{[1,s]}\\ \tau\in\{0,1,-1\}^{[1,r]}\end{subarray}}\left(\bigwedge_{j=1}^{s}(\mbox{\bf sign}(h_{j})=\sigma(j))\wedge\bigwedge_{i=1}^{r}(\mbox{\bf sign}(g_{i})=\tau(i))\right)\Longrightarrow\\ \left(T-\sum_{j=1}^{s}\sigma(j)h_{j}-\frac{1}{2}\sum_{i=1}^{r}\tau(i)(1+\tau(i))g_{i}=0\right)

describes the graph of ψ\psi and all polynomials occurring in it have degrees at most dd.

Moreover, it follows from Lemma 5.3 that the graph of dist(,M)\operatorname{dist}(\cdot,M) can be defined by a quantifier-free formula involving polynomials in n+1n+1 variables having degrees bounded by dO(n)d^{O(n)}. The first part of the theorem now follows from Theorem 1 after setting f(x)=ψ(x)f(x)=\psi(x), and g(x)=dist(x,M)g(x)=\operatorname{dist}(x,M).

In case that the MM is finite, it is possible to derive a sharper estimate of the error bound exponent. Suppose M={p1,,pN}RnM=\{p_{1},\ldots,p_{N}\}\subset\mathrm{R}^{n}. In this case the graph of the distance function, dist(X,M)\operatorname{dist}(X,M), is described by the following formula

Θ(X,T):=(T0)(i=1N(T2Xpi2))(i=1N(T2=Xpi2)).\Theta(X,T):=(T\geq 0)\wedge\left(\bigwedge_{i=1}^{N}(T^{2}\geq||X-p_{i}||^{2})\right)\wedge\left(\bigvee_{i=1}^{N}(T^{2}=||X-p_{i}||^{2})\right).

Notice that the degrees of the polynomials appearing in the quantifier-free formula Θ\Theta are bounded by 22. Furthermore, the graph of the residual function ψ\psi is defined by quantifier-free formula involving polynomials of degrees bounded by dd. The second part of the theorem is now immediate from Theorem 1. ∎

Remark 9.

A weaker version of Theorem 2 can be directly obtained using the quantifier elimination. Let us define

Mε:={xRnψ(x)=ε,xE}.\displaystyle M_{\varepsilon}:=\{x\in\mathrm{R}^{n}\mid\psi(x)=\varepsilon,\ x\in E\}.

Then it suffices to describe the graph of the semi-algebraic function

φ:εsupxMεdist(x,M).\displaystyle\varphi:\varepsilon\mapsto\sup_{x\in M_{\varepsilon}}\operatorname{dist}(x,M).

Let Φ(X,ε)\Phi(X,\varepsilon) be a quantifier-free formula describing MεM_{\varepsilon} and Ψ(X,T)\Psi(X,T) be the quantifier-free formula describing the graph of dist(x,M)\operatorname{dist}(x,M) in the proof of Lemma 5.3. Then the graph of φ\varphi is the lower envelope of the realization of the formula (5.5):

(5.5) (X)(T)¬(Φ(X,ε)Ψ(X,T))(φT).\displaystyle(\forall X)\ (\exists T)\quad\neg(\Phi(X,\varepsilon)\wedge\Psi(X,T))\vee(\varphi\geq T).

By Lemma 5.3 and the application of Theorem 4 to (5.5), the graph of φ\varphi can be described by a quantifier-free 𝒫\mathcal{P}-formula with 𝒫R[φ,ε]dO(n2)\mathcal{P}\subset\mathrm{R}[\varphi,\varepsilon]_{\leq d^{O(n^{2})}}, which implies the existence of P𝒫P\in\mathcal{P} such that P(φ,ε)=0P(\varphi,\varepsilon)=0 for sufficiently small positive ε\varepsilon. By the Newton-Puiseux theorem [48, Theorems 3.2 and 4.1 of Chapter IV], ψ\psi near ε=0\varepsilon=0 can be expanded as

φ(ε)=a1εγ1+a2εγ2+\displaystyle\varphi(\varepsilon)=a_{1}\varepsilon^{\gamma_{1}}+a_{2}\varepsilon^{\gamma_{2}}+\ldots

where 0<γ1<γ2<0<\gamma_{1}<\gamma_{2}<\ldots, and aiRa_{i}\in\mathrm{R}. Furthermore, it follows from the Newton-Puiseux algorithm [48, Section 3.2 of Chapter IV] that γ11/deg(P)\gamma_{1}\leq 1/\deg(P), where deg(P)=dO(n2)\deg(P)=d^{O(n^{2})}. All this implies φdeg(P)=O(ε)\varphi^{\deg(P)}=O(\varepsilon) for sufficiently small positive ε\varepsilon, and thus

dist(x,M)deg(P)κψ(x),for all xE with sufficiently small ψ(x).\displaystyle\operatorname{dist}(x,M)^{\deg(P)}\leq\kappa\cdot\psi(x),\quad\mbox{for all }x\in E\mbox{ with sufficiently small }\psi(x).

The case with dimM=0\dim M=0 is analogous.

Proof of Corollary 1.

Notice that M𝕊+pM\cap\mathbb{S}^{p}_{+} can be redefined as a basic 𝒬\mathcal{Q}-semi-algebraic set with 𝒬R[X1,,Xp2]max{d,p}\mathcal{Q}\subset\mathrm{R}[X_{1},\ldots,X_{p^{2}}]_{\leq\max\{d,p\}} and |𝒬|=2p+r1|\mathcal{Q}|=2^{p}+r-1 as follows:

{X𝕊pgi(X)0,i=1,,r,det(XI)0,I{1,,p}},\displaystyle\big{\{}X\in\mathbb{S}^{p}\mid g_{i}(X)\leq 0,\ i=1,\ldots,r,\ \det(X_{I})\geq 0,\ \forall I\subseteq\{1,\ldots,p\}\big{\}},

where XIX_{I} is a principal submatrix of XX indexed by II. We also define the residual function

ψ(x):=max{dist(x,M),maxI{1,,p}(max{det(XI),0})},\displaystyle\psi(x):=\max\Big{\{}\operatorname{dist}(x,M),\max_{I\subseteq\{1,\ldots,p\}}\big{(}\max\{-\det(X_{I}),0\}\big{)}\Big{\}},

which is a 𝒬\mathcal{Q}^{\prime}-semi-algebraic function with 𝒬R[X1,,Xp2,Y]dO(p2)\mathcal{Q}^{\prime}\subset\mathrm{R}[X_{1},\ldots,X_{p^{2}},Y]_{\leq d^{O(p^{2})}}, see Lemma 5.3 and the proof of Theorem 2. Then by applying Theorem 2, Lemma 5.3, and Remark 6 to M𝕊+pM\cap\mathbb{S}^{p}_{+}, EE, and ψ(x)\psi(x) we get

dist(x,M𝕊+p)ρκmax{dist(x,M),maxI{1,,p}(max{det(XI),0})}, for all xE,\operatorname{dist}(x,M\cap\mathbb{S}_{+}^{p})^{\rho}\leq\kappa^{\prime}\cdot\max\Big{\{}\operatorname{dist}(x,M),\max_{I\subseteq\{1,\ldots,p\}}\big{(}\max\{-\det(X_{I}),0\}\big{)}\Big{\}},\\ \mbox{ for all }x\in E,

for some κ>0\kappa^{\prime}>0 and ρ=max{d,p}O(p4)\rho=\max\{d,p\}^{O(p^{4})}. The rest of the proof follows from the arguments in [28]. Let λi(XI)\lambda_{i}(X_{I}) be the iith smallest eigenvalue of XIX_{I}. By the boundedness of EE, there exists some positive rr such that |λi(XI)|r|\lambda_{i}(X_{I})|\leq r for all i=1,,|I|i=1,\ldots,|I|. Furthermore, by the interlacing property of eigenvalues of XX, we have

λmin(XI):=λ1(XI)λmin(X),\displaystyle\lambda_{\min}(X_{I}):=\lambda_{1}(X_{I})\geq\lambda_{\min}(X),

which yields

det(XI)=i=1|I|λi(XI)r|I|1λmin(XI)r|I|1λmin(X).\displaystyle\det(X_{I})=\prod_{i=1}^{|I|}\lambda_{i}(X_{I})\geq r^{|I|-1}\lambda_{\min}(X_{I})\geq r^{|I|-1}\lambda_{\min}(X).

Consequently,

max{det(XI),0}r|I|1max{λmin(X),0}\displaystyle\max\{-\det(X_{I}),0\}\leq r^{|I|-1}\cdot\max\{-\lambda_{\min}(X),0\}

for every I{1,,p}I\subseteq\{1,\ldots,p\}, which completes the proof. ∎

5.3. Proof of Theorem 3

In the proof of Theorem 3 we need the following ingredient which is proved in [3, Theorem 2.5].

Proposition 5.4 (Quantitative cylindrical definable cell decomposition).

Fix an o-minimal expansion of the real closed field R\mathrm{R}. Let 𝒜\mathcal{A} be a definable family of subsets of Rn\mathrm{R}^{n} parametrized by the definable set AA. Then there exist a finite set JJ and definable families (𝒞j)jJ(\mathcal{C}_{j})_{j\in J} of subsets of Rn\mathrm{R}^{n} each parametrized by AN(n)A^{N(n)} where N(n)=2(2n1)N(n)=2(2^{n}-1) having the following property. Suppose, AAA^{\prime}\subset A is a finite subset. Then, there exists a cylindrical decomposition 𝒟=(𝒟1,,𝒟n)\mathcal{D}=(\mathcal{D}_{1},\ldots,\mathcal{D}_{n}) of Rn\mathrm{R}^{n}, such that for each i,1ini,1\leq i\leq n, the set of cells of the cylindrical decomposition (𝒟1,,𝒟i)(\mathcal{D}_{1},\ldots,\mathcal{D}_{i}) of Ri\mathrm{R}^{i} is a subset of the set of definable sets

{(𝒞j)a¯a¯AN(n)}.\{(\mathcal{C}_{j})_{\overline{a}}\;\mid\;\overline{a}\in A^{\prime N(n)}\}.

For the rest of this section we fix a polynomially bounded o-minimal expansion of \mathbb{R}.

Proposition 5.5 (o-minimal quantitative version of Proposition 2.6.4 in [10]).

Let 𝒜\mathcal{A} be a definable family of subsets of n\mathbb{R}^{n} parametrized by the definable set AA, and let \mathcal{B} be a definable subset of n+1\mathbb{R}^{n+1} parametrized by the definable set BB.

Then there exists N=N(𝒜,)>0N=N(\mathcal{A},\mathcal{B})>0 having the following property. For any triple of finite sets (A,B,B′′)(A^{\prime},B^{\prime},B^{\prime\prime}) with AA,B,B′′BA^{\prime}\subset A,B^{\prime},B^{\prime\prime}\subset B, a closed (𝒜,A)(\mathcal{A},A^{\prime})-set SS, a (,B)(\mathcal{B},B^{\prime})-set FF, (,B′′)(\mathcal{B},B^{\prime\prime})-set GG, such that F,GF,G are graphs of definable functions f:n,g:nf:\mathbb{R}^{n}\rightarrow\mathbb{R},g:\mathbb{R}^{n}\rightarrow\mathbb{R}, such that f|Sf|_{S} and g|Sf0g|_{S_{f\neq 0}} (where Sf0={xSf(x)0}S_{f\neq 0}=\{x\in S\;\mid\;f(x)\neq 0\}) are continuous, the function fNg|Sf0f^{N}g|_{S_{f\neq 0}} extended by 0 on {xSf(x)=0}\{x\in S\;\mid\;f(x)=0\} is continuous on SS.

Proof.

We follow closely the proof of the corresponding proposition (Proposition 5.2) in the semi-algebraic case.

For each xS,ux\in S,u\in\mathbb{R}, we define

Sx,u={ySyx1,u|f(y)|=1}.S_{x,u}=\{y\in S\;\mid\;||y-x||\leq 1,u|f(y)|=1\}.

The set Sx,uS_{x,u} is definable, closed and bounded, and we define the function

v(x,u)={0 if Ax,u=,sup{|g(y)|yAx,u} otherwise.v(x,u)=\begin{cases}0\mbox{ if }A_{x,u}=\emptyset,\\ \sup\{|g(y)|\;\mid\;y\in A_{x,u}\}\mbox{ otherwise}.\end{cases}

Clearly, v:S×v:S\times\mathbb{R}\rightarrow\mathbb{R} is a definable function.

We define

 T={(x,u,y,v)yS(||yx||210)    ((v>0(uv1=0))((v<0)(uv+1=0)))    (y,v)F}. \halign{\hbox to\displaywidth{$\hfil\displaystyle#\hfil$}\cr 0.0pt{\hfil$\displaystyle T=\{(x,u,y,v)\mid y\in S\wedge(||y-x||^{2}-1\leq 0)\cr 0.0pt{\hfil$\displaystyle\;\wedge\;\cr 0.0pt{\hfil$\displaystyle((v>0\;\wedge\;(uv-1=0))\;\vee\;((v<0)\;\wedge\;(uv+1=0)))\cr 0.0pt{\hfil$\displaystyle\;\wedge\;\cr 0.0pt{\hfil$\displaystyle(y,v)\in F\}.\crcr}}}}}}

Notice that TT is definable and for each fixed xx, TxRn+2T_{x}\subset\mathrm{R}^{n+2} is a (𝒞,A×B)(\mathcal{C},A^{\prime}\times B^{\prime})-set where 𝒞\mathcal{C} is a definable family of subsets of Rn+2\mathrm{R}^{n+2} parametrized by (A×B)(A\times B), and 𝒞\mathcal{C} depends only on the definable families 𝒜,\mathcal{A},\mathcal{B}.

Observe also that for each xAx\in A and uRu\in\mathrm{R},

Tx,u=graph(f|Sx,u).T_{x,u}=\mathrm{graph}(f|_{S_{x,u}}).

We also define the set L0L_{0} by

 L0={(x,u,w)(y,v,z)  (w0)    ((x,u,y,v)T(y,z)G)  ((z0)(wz))((z0)(wz)))}. \halign{\hbox to\displaywidth{$\hfil\displaystyle#\hfil$}\cr 0.0pt{\hfil$\displaystyle L_{0}=\{(x,u,w)\;\mid\;\forall(y,v,z)\cr 0.0pt{\hfil$\displaystyle(w\geq 0)\cr 0.0pt{\hfil$\displaystyle\wedge\cr 0.0pt{\hfil$\displaystyle((x,u,y,v)\in T\wedge(y,z)\in G)\Longrightarrow\cr 0.0pt{\hfil$\displaystyle((z\geq 0)\wedge(w\geq z))\vee((z\leq 0)\wedge(w\geq-z)))\}.\crcr}}}}}}

Finally, let

L={(x,u,w)(w)(x,u,w)L0(0ww)}.L=\{(x,u,w)\;\mid\;(\forall w^{\prime})(x,u,w^{\prime})\in L_{0}\Longrightarrow(0\leq w\leq w^{\prime})\}.

Notice that for xSx\in S, (x,u,w)L(x,u,w)\in L if and only if w=v(x,u)w=v(x,u).

Also notice that the set LS×2L\subset S\times\mathbb{R}^{2} is a definable set which is the complement of a projection of an (𝒟,D)(\mathcal{D},D^{\prime})-set P2n+5P\subset\mathbb{R}^{2n+5}, where 𝒟\mathcal{D} is a definable family of sets parametrized by A×A×B×BA\times A\times B\times B and D=A×A×B×B′′D^{\prime}=A^{\prime}\times A^{\prime}\times B^{\prime}\times B^{\prime\prime}, and 𝒟\mathcal{D} depends only on the definable families 𝒜,\mathcal{A},\mathcal{B}.

We now apply Proposition 5.4 to the definable family 𝒟\mathcal{D}. There exist a set of definable families (𝒞j)jJ(\mathcal{C}_{j})_{j\in J} and a cylindrical decomposition (𝒟1,,𝒟2n+5)(\mathcal{D}_{1},\ldots,\mathcal{D}_{2n+5}) of 2n+5\mathbb{R}^{2n+5} adapted to the set PP whose cells are of the form (𝒞j)w(\mathcal{C}_{j})_{w} with jJ,w(D)N(2n+5)j\in J,w\in(D^{\prime})^{N(2n+5)}.

For xSx\in S, there exists a unique cell, C=(𝒞j)wC=(\mathcal{C}_{j})_{w}, jJ,w(D)N(2n+5)j\in J,w\in(D^{\prime})^{N(2n+5)}, of the decomposition n\mathbb{R}^{n} containing xx.

The definable functions v(x,)v(x,\cdot) as xx varies over (𝒞j)w(\mathcal{C}_{j})_{w} and ww varies over (A×A×B×B)N(2n+5)(A\times A\times B\times B)^{N(2n+5)} and jJj\in J form a definable family, and using Proposition 5.2 in [38] there exists p=p(𝒜,)p=p(\mathcal{A},\mathcal{B}) such that

|v(x,u)|c(x)up,|v(x,u)|\leq c(x)\cdot u^{p},

for all uu large enough.

Now for each xSx\in S, there exists a cell, C=(𝒞j)wC=(\mathcal{C}_{j})_{w} jJ,w(D)N(2n+5)j\in J,w\in(D^{\prime})^{N(2n+5)}, of the decomposition n\mathbb{R}^{n} containing xx.

This proves the proposition taking N=p+1N=p+1. ∎

Theorem 6.

Let 𝒜\mathcal{A} be a definable family of subsets of Rn\mathrm{R}^{n} parametrized by the definable set AA, and let \mathcal{B} be a definable subset of Rn+1\mathrm{R}^{n+1} parametrized by the definable set BB. Then there exists N=N(𝒜,)>0N=N(\mathcal{A},\mathcal{B})>0 having the following property. For any triple of finite sets (A,B,B′′)(A^{\prime},B^{\prime},B^{\prime\prime}) with AA,B,B′′BA^{\prime}\subset A,B^{\prime},B^{\prime\prime}\subset B, a closed (𝒜,A)(\mathcal{A},A^{\prime})-set SS, a (,B)(\mathcal{B},B^{\prime})-set FF, (,B′′)(\mathcal{B},B^{\prime\prime})-set GG, such that F,GF,G are graphs of definable functions f:n,g:nf:\mathbb{R}^{n}\rightarrow\mathbb{R},g:\mathbb{R}^{n}\rightarrow\mathbb{R} continuous on SS, such that f|S1(0)g|S1(0)f|_{S}^{-1}(0)\subset g|_{S}^{-1}(0). Then there exists a continuous definable function h:SRh:S\rightarrow\mathrm{R} such that g|SN=hf|Sg|_{S}^{N}=hf|_{S} on SS.

Proof.

Similar to the proof of Theorem 5 replacing semi-algebraic by definable. ∎

Proof of Theorem 3.

Use Theorem 6 with c=supxS|h(x)|c=\sup_{x\in S}|h(x)|. ∎

6. Applications to optimization

As an illustration of the improvement one obtains by applying the improved bound on the Łojasiewicz exponent proved in Theorem 1 and the error bound in Theorem 2 we consider the following application in the theory of optimization. Clearly, Theorem 2 can be applied to other situations as well where error bounds are important, for example in the study of Hölderian continuity of the set-valued map defined by (2.5) as stated in [37, Theorem 3.1].

6.1. Binary feasibility problems

We can use Theorem 2 and its independence from the number of constraints to derive an error bound for a binary feasibility problem, where the feasible set is defined by

(6.1) M={xRngi(x)0,hj(x)\displaystyle M=\{x\in\mathrm{R}^{n}\mid g_{i}(x)\leq 0,\ h_{j}(x) =0,i=1,,r,j=1,,s,\displaystyle=0,\ i=1,\ldots,r,\ j=1,\ldots,s,
xk\displaystyle x_{k} {0,1},k=1,,n},\displaystyle\in\{0,1\},\ k=1,\ldots,n\},

where gi,hjR[X1,,Xn]dg_{i},h_{j}\in\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d}. The following result is a quantified version of [37, Theorem 5.6] specialized for polynomials.

Corollary 2.

Let MM be defined in (6.1) and let EE be a closed and bounded 𝒫\mathcal{P}-semi-algebraic subset of Rn\mathrm{R}^{n} with 𝒫R[X1,,Xn]d\mathcal{P}\subset\mathrm{R}[X_{1},\ldots,X_{n}]_{\leq d}. Then there exist κ>0\kappa>0 and ρ=(O(d))2n+14\rho=(O(d))^{2n+14} such that

dist(x,M)ρκψ(x),for all xE,\displaystyle\operatorname{dist}(x,M)^{\rho}\leq\kappa\cdot\psi(x),\quad\mbox{for all }x\in E,

where

ψ(x)=j=1s(hj(x))2+i=1r(max{gi(x),0})2+k=1n|xk(1xk)|.\displaystyle\psi(x)=\sqrt{\sum_{j=1}^{s}(h_{j}(x))^{2}}+\sqrt{\sum_{i=1}^{r}(\max\{g_{i}(x),0\})^{2}}+\sum_{k=1}^{n}\lvert x_{k}(1-x_{k})\rvert.
Proof.

Note that for every k=1,,nk=1,\ldots,n, the binary constraint on xkx_{k} can be enforced by

xk(1xk)=0,xk[0,1]R.\displaystyle x_{k}(1-x_{k})=0,\quad x_{k}\in[0,1]\subset\mathrm{R}.

Then MM can be redefined as

(6.2) M:={xRngi(x)0,hj(x)\displaystyle M:=\Big{\{}x\in\mathrm{R}^{n}\mid g_{i}(x)\leq 0,\ h_{j}(x) =0,\displaystyle=0, i\displaystyle i =1,,r,j=1,,s,\displaystyle=1,\ldots,r,\ j=1,\ldots,s,
xk(1xk)\displaystyle x_{k}(1-x_{k}) =0,\displaystyle=0, k\displaystyle k =1,,n,},\displaystyle=1,\ldots,n,\Big{\}},

which is a finite subset of Rn\mathrm{R}^{n}.

Also observe that defining

ψ~(x)=j=1s|hj(x)|+i=1rmax{gi(x),0}+k=1n|xk(1xk)|,\displaystyle\widetilde{\psi}(x)=\sum_{j=1}^{s}|h_{j}(x)|+\sum_{i=1}^{r}\max\{g_{i}(x),0\}+\sum_{k=1}^{n}\lvert x_{k}(1-x_{k})\rvert,

it follows from the Cauchy-Schwarz inequality that

ψ~(x)cψ(x),\widetilde{\psi}(x)\leq c\cdot\psi(x),

with c=max{1,r,s}>0c=\max\{1,\sqrt{r},\sqrt{s}\}>0.

Now the result follows by applying the second part of Theorem 2 and Remark 6 to (6.2) and the residual function ψ~(x)\widetilde{\psi}(x). ∎

6.2. Convergence rate of feasible descent schemes

Error bounds are important to estimate the convergence rate of iterative algorithms in nonlinear optimization. Here, we present the convergence analysis in [36, Theorem 5], which is relevant to Theorem 1.

Let R=\mathrm{R}=\mathbb{R} and gi,hjg_{i},h_{j} defined in (2.5) be convex polynomials. Then the goal of a feasible descent scheme is to find stationary solutions of a polynomial f[X1,,Xn]df\in\mathbb{R}[X_{1},\ldots,X_{n}]_{\leq d} over a convex MM (assuming that infxMf>\inf_{x\in M}f>-\infty), where a stationary solution is an xnx\in\mathbb{R}^{n} such that

xProjM(xf(x))=0,\displaystyle x-\operatorname{Proj}_{M}\!\big{(}x-\nabla f(x)\big{)}=0,

in which ProjM()\operatorname{Proj}_{M}(\cdot) denotes the projection onto the convex set MM. The idea of a feasible descent scheme is to generate a sequence {xk}k=1\{x_{k}\}_{k=1}^{\infty} of solutions by

(6.3) xk+1=ProjM(xkαkf(xk)+ek),\displaystyle x_{k+1}=\operatorname{Proj}_{M}\!\big{(}x_{k}-\alpha_{k}\nabla f(x_{k})+e_{k}\big{)},

where αk>0\alpha_{k}>0 is so-called the step length and eke_{k} is an error vector depending on xkx_{k}. If we define the set of stationary solutions as

S:={xnxProjM(xf(x))=0},\displaystyle S:=\Big{\{}x\in\mathbb{R}^{n}\mid x-\operatorname{Proj}_{M}\!\big{(}x-\nabla f(x)\big{)}=0\Big{\}},

then the following result is well-known for the convergence rate of a feasible descent scheme which we specialize for the polynomial ff.

Proposition 6.1 (Theorem 5 in [36]).

Suppose that the gradient of ff is Lipschitz continuous and there exists ε>0\varepsilon>0 such that

xS,yS,f(x)f(y)xyε.\displaystyle x\in S,\ \ y\in S,\ \ f(x)\neq f(y)\Longrightarrow\|x-y\|\geq\varepsilon.

If lim infαk>0\liminf\alpha_{k}>0 and the sequences {ek}k=1\{e_{k}\}_{k=1}^{\infty} and {xk}k=1\{x_{k}\}_{k=1}^{\infty} generated by (6.3) satisfy

ekκ1xkxk+1,for someκ1>0,\displaystyle\|e_{k}\|\leq\kappa_{1}\|x_{k}-x_{k+1}\|,\ \text{for some}\ \kappa_{1}>0,
f(xk+1)f(xk)κ2xkxk+12,for someκ2>0,\displaystyle f(x_{k+1})-f(x_{k})\leq-\kappa_{2}\|x_{k}-x_{k+1}\|^{2},\ \text{for some}\ \kappa_{2}>0,

then the sequence {f(xk)}k=1\{f(x_{k})\}_{k=1}^{\infty} converges at least sub-linearly, at the rate k1ρk^{1-\rho}, where ρ1\rho\geq 1 is an integer satisfying

dist(x,S)ρκxProjM(xf(x))\displaystyle\operatorname{dist}(x,S)^{\rho}\leq\kappa\cdot\|x-\operatorname{Proj}_{M}\!\big{(}x-\nabla f(x)\big{)}\|

for some κ>0\kappa>0 and for all xx in a compact semi-algebraic subset of MM.

Notice that SS and xProjM(xf(x))\|x-\operatorname{Proj}_{M}\!\big{(}x-\nabla f(x)\big{)}\| are both semi-algebraic, and the latter is a residual function. Therefore, Theorem 1 and Lemma 5.3 can be applied to quantify the convergence rate k1ρk^{1-\rho} in terms of dd and nn only.

Remark 10.

The upper bound (3.6) was used to quantify the convergence rate of the cyclic projection algorithm applied to finite intersections of convex semi-algebraic subsets of n\mathbb{R}^{n}  [11, Theorem 4.4].

6.3. Sums of squares relaxation

A polynomial optimization problem is formally defined as the following: given a polynomial f[X1,,Xn]df\in\mathbb{R}[X_{1},\ldots,X_{n}]_{\leq d}, compute

(6.4) fmin:=inf{f(x)xM},\displaystyle f_{\min}^{*}:=\inf\big{\{}f(x)\mid x\in M\big{\}},

where MM is defined in (2.5) with R=\mathrm{R}=\mathbb{R}. We assume, without loss of generality, here that s=0s=0 in (2.5).

Unlike semi-definite optimization, see e.g., [5], there is no efficient interior point method for polynomial optimization. Nevertheless, tools in real algebra have laid the groundwork for developing an efficient numerical approach, where semi-definite relaxation plays a central role. Using this approach, (6.4) is approximated by a hierarchy of semi-definite relaxations, so-called sums of squares (SOS) relaxations [26, 44], see also [27]. A SOS relaxation of order tt is defined as

(6.5) fsost=sup{βfβ2t(g1,,gr)},\displaystyle f_{\mathrm{sos}}^{*t}=\sup\{\beta\mid f-\beta\in\mathbf{\mathcal{M}}_{2t}(g_{1},\ldots,g_{r})\},

where

2t(g1,,gr)\displaystyle\mathbf{\mathcal{M}}_{2t}(g_{1},\ldots,g_{r}) =\displaystyle=
{u0j=1rujgju0,ujΣ,deg(u0),deg(ujgj)2t,j=1,,r}\displaystyle\bigg{\{}u_{0}-\sum_{j=1}^{r}u_{j}g_{j}\mid u_{0},u_{j}\in\Sigma,\ \deg(u_{0}),\deg(u_{j}g_{j})\leq 2t,\ j=1,\ldots,r\bigg{\}}

is called the truncated quadratic module generated by g1,,grg_{1},\ldots,g_{r} and Σ\Sigma is the convex cone of SOS polynomials. It is worth noting that (6.5) is a semi-definite optimization problem of size O(nt)O(n^{t}). Under some conditions on MM, see e.g., [27, Proposition 6.2 and Theorem 6.8], (6.5) is feasible for sufficiently large tt, and fsostfminf_{\mathrm{sos}}^{*t}\to f_{\min}^{*} as tt\to\infty.

Recently, Baldi and Mourrain [1] provided a convergence rate for fsostf_{\mathrm{sos}}^{*t} in terms of tt, in which the Łojasiewicz exponent plays a central role. The authors proved [1, Theorem 4.3], under some conditions, that there exists c>1c>1 such that

0<fminfsostcfdeg(f)75t12.5nρ,\displaystyle 0<f_{\min}^{*}-f_{\mathrm{sos}}^{*t}\leq c\cdot\|f\|\deg(f)^{\frac{7}{5}}t^{-\frac{1}{2.5n\rho}},

where cc depends on nn, ρ\rho, and g1,,grg_{1},\ldots,g_{r}, f:=maxx[1,1]n|f(x)|\|f\|:=\max_{x\in[-1,1]^{n}}|f(x)|, and ρ\rho is the error bound exponent for the inequality

dist(x,M)ρκ|min{g1(x),,gr(x),0}|,for allx[0,1]n\displaystyle\operatorname{dist}(x,M)^{\rho}\leq\kappa\cdot|\min\{g_{1}(x),\ldots,g_{r}(x),0\}|,\quad\text{for all}\ x\in[0,1]^{n}

for some κ>0\kappa>0. Then the application of Theorem 2 and Remark 6 yields an upper bound dO(n2)d^{O(n^{2})} on ρ\rho and thus proves a lower bound on the convergence rate t12.5nρt^{-\frac{1}{2.5n\rho}} of the SOS relaxation in terms of nn and dd only.

Remark 11.

There are other applications of the Lojasiewicz inequality to polynomial optimization in the literature. For instance, it was shown in [23] that the Łojasiewicz inequality (3.2) can be used to reduce (6.4) to minimization over a ball.

7. Conclusion

In this paper, we proved a nearly tight upper bound on the Łojasiewicz exponent for semi-algebraic functions over a real closed field R\mathrm{R} in a very general setting. Unlike the previous best known bound in this setting due to Solernó [45], our bound is independent of the cardinalities of the semi-algebraic descriptions of ff, gg, and AA. We exploited this fact to improve the best known error bounds for polynomial and non-linear semi-definite systems. As an abstraction of the notion of independence from the combinatorial parameters, we proved a version of Łojasiewicz inequality in polynomially bounded o-minimal structures. We proved existence of a common Łojasiewicz exponent for certain combinatorially defined infinite (but not necessarily definable) families of pairs of functions, which improves a prior result due to Chris Miller [38].

We end with a few open problems. We proved in Theorem 2 that the exponent ρ=dO(n)\rho=d^{O(n)} in the error bound with respect to a zero-dimensional semi-algebraic set MM, and Example 1 indicates that this bound is indeed tight. Without the assumption on the dimension on MM, the general bound on the exponent ρ\rho in Theorem 2 is dO(n2)d^{O(n^{2})}. There are some indications in [21] for generating examples whose Łojasiewicz exponent is worse than Example 1. However, we have not been able to find any example with ρ=dO(n2)\rho=d^{O(n^{2})}, so we do not know if this bound is tight as well. It would be interesting to resolve this gap.

Another interesting question is to prove an upper bound that depends on dimM\dim M which interpolates between the 0-dimensional and the general case. More precisely, is it possible to improve the upper bound in Theorem 2 to dO(ndimM)d^{O(n\cdot\dim M)} ?

While the emphasis in the current paper has been on proving a bound on the Łojasiewicz exponent which is independent of the combinatorial parameter there is a special case that merits attention and in which the combinatorial parameter may play a role. It is well known [2] that the topological complexity (say measured in terms of the Betti numbers) of a real algebraic set in Rn\mathrm{R}^{n} defined by ss quadratic equations is bounded by nO(s)n^{O(s)}. This bound (unlike the bounds discussed in the current paper) is polynomial in nn for fixed ss. One could ask if a similar bound also holds in this setting for the Łojasiewicz exponent.

Acknowledgements

The first author was partially supported by NSF grants CCF-1910441 and CCF-2128702. The second author was partially supported by NSF grant CCF-2128702.

References

  • [1] L. Baldi and B. Mourrain, On the effective Putinar’s positivstellensatz and moment approximation, (2022). arXiv:2111.11258v2 https://arxiv.org/abs/2111.11258.
  • [2] A. I. Barvinok, On the Betti numbers of semi-algebraic sets defined by few quadratic inequalities, Math. Z., 225 (1997), pp. 231–244.
  • [3] S. Basu, Combinatorial complexity in o-minimal geometry, Proc. London Math. Soc. (3), 100 (2010), pp. 405–428.
  • [4] S. Basu, Algorithms in real algebraic geometry: a survey, in Real algebraic geometry, vol. 51 of Panor. Synthèses, Soc. Math. France, Paris, 2017, pp. 107–153.
  • [5] S. Basu and A. Mohammad-Nezhad, On the central path of semi-definite optimization: Degree and worst-case convergence rate, SIAM Journal on Applied Algebra and Geometry, 6 (2022), pp. 299–318.
  • [6] S. Basu, R. Pollack, and M.-F. Roy, On the combinatorial and algebraic complexity of quantifier elimination, J. ACM, 43 (1996), pp. 1002–1045.
  • [7]  , Algorithms in real algebraic geometry, vol. 10 of Algorithms and Computation in Mathematics, Springer-Verlag, Berlin, version bpr-ed2-posted3 from 27/06/2016 available at http://perso.univ-rennes1.fr/marie-francoise.roy/bpr-ed2-posted3.html.
  • [8] S. Basu and M.-F. Roy, Bounding the radii of balls meeting every connected component of semi-algebraic sets, Journal of Symbolic Computation, 45 (2010), pp. 1270–1279.
  • [9] S. Basu and M.-F. Roy, Quantitative curve selection lemma, Mathematische Zeitschrift, (2021).
  • [10] J. Bochnak, M. Coste, and M.-F. Roy, Real Algebraic Geometry, Springer, 1998.
  • [11] J. M. Borwein, G. Li, and L. Yao, Analysis of the convergence rate for the cyclic projection algorithm applied to basic semi-algebraic convex sets, SIAM Journal on Optimization, 24 (2014), pp. 498–527.
  • [12] L. Bröcker, On basic semi-algebraic sets, Expositiones Mathematicae, 9 (1991), pp. 289–334.
  • [13] M. Coste, An introduction to o-minimal geometry, Istituti Editoriali e Poligrafici Internazionali, Pisa, 2000. Dip. Mat. Univ. Pisa, Dottorato di Ricerca in Matematica.
  • [14] D. D’Acunto and K. Kurdyka, Explicit bounds for the Łojasiewicz exponent in the gradient inequality for polynomials, Annales Polonici Mathematici, 87 (2005), pp. 51–61.
  • [15] E. de Klerk, Aspects of Semi-definite Programming: Interior Point Algorithms and Selected Applications, vol. 65 of Series Applied Optimization, Springer, 2006.
  • [16] A. Gabrielov and N. Vorobjov, Approximation of definable sets by compact families, and upper bounds on homotopy and homology, J. Lond. Math. Soc. (2), 80 (2009), pp. 35–54.
  • [17] J. Gwoździewicz, The Łojasiewicz exponent of an analytic function at an isolated zero, Commentarii Mathematici Helvetici, 74 (1999), pp. 364–375.
  • [18] L. Hörmander, On the division of distributions by polynomials, Ark. Mat., 3 (1958), pp. 555–568.
  • [19] L. Hörmander, On the division of distributions by polynomials, Arkiv för Matematik, 3 (1958), pp. 555–568.
  • [20] S. Ji, J. Kollar, and B. Shiffman, A global Łojasiewicz inequality for algebraic varieties, Transactions of the American Mathematical Society, 329 (1992), pp. 813–818.
  • [21] J. Kollár, An effective Łojasiewicz inequality for real polynomials, Periodica Mathematica Hungarica, 38 (1999), pp. 213–221.
  • [22] K. Kurdyka and S. Spodzieja, Separation of real algebraic sets and the Łojasiewicz exponent, Proceedings of the American Mathematical Society., 142 (2014-9-14), pp. 3089–3102.
  • [23]  , Convexifying positive polynomials and sums of squares approximation, SIAM Journal on Optimization, 25 (2015), pp. 2512–2536.
  • [24] K. Kurdyka, S. Spodzieja, and A. Szlachcińska, Metric properties of semi-algebraic mappings, Discrete & Computational Geometry, 55 (2016), pp. 786–800.
  • [25]  , Correction to: Metric properties of semi-algebraic mappings, Discrete & Computational Geometry, 62 (2019), pp. 990–991.
  • [26] J. B. Lasserre, Global optimization with polynomials and the problem of moments, SIAM Journal on Optimization, 11 (2001), pp. 796–817.
  • [27] M. Laurent, Sums of Squares, Moment Matrices and Optimization Over Polynomials, Springer New York, New York, NY, 2009, pp. 157–270.
  • [28] A. S. Lewis and J.-S. Pang, Error Bounds for Convex Inequality Systems, Springer US, Boston, MA, 1998, pp. 75–110.
  • [29] G. Li, On the asymptotically well behaved functions and global error bound for convex polynomials, SIAM Journal on Optimization, 20 (2010), pp. 1923–1943.
  • [30] G. Li, B. S. Mordukhovich, and T. S. Pham, New fractional error bounds for polynomial systems with applications to hölderian stability in optimization and spectral theory of tensors, Mathematical Programming, 153 (2015), pp. 333–362.
  • [31] S. Łojasiewicz, Sur le problème de la division, Studia Math., 18 (1959), pp. 87–136.
  • [32] S. Łojasiewicz, Triangulation of semi-analytic sets., Ann. Scuola Norm. Sup. Pisa, Sci. Fis. Mat., 18 (1964), pp. 449–474.
  • [33]  , Ensembles semi-analytiques. preprint, IHES, 1965.
  • [34] S. Łojasiewicz, Sur les ensembles semi-analytiques, (1971), pp. 237–241.
  • [35] X.-D. Luo and Z.-Q. Luo, Extension of Hoffman’s error bound to polynomial systems, SIAM Journal on Optimization, 4 (1994), pp. 383–392.
  • [36] Z.-Q. Luo, New error bounds and their applications to convergence analysis of iterative algorithms, Mathematical Programming, 88 (2000), pp. 341–355.
  • [37] Z.-Q. Luo and J.-S. Pang, Error bounds for analytic systems and their applications, Mathematical Programming, 67 (1994), pp. 1–28.
  • [38] C. Miller, Expansions of the real field with power functions, Ann. Pure Appl. Logic, 68 (1994), pp. 79–94.
  • [39] J. Milnor, Singular points of complex hypersurfaces, Annals of Mathematics Studies, No. 61, Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1968.
  • [40] L. M. G. na Drummond and Y. Peterzil, The central path in smooth convex semidefinite programs, Optimization, 51 (2002), pp. 207–233.
  • [41] B. Osińska-Ulrych, G. Skalski, and S. Spodzieja, Effective Łojasiewicz gradient inequality for Nash functions with application to finite determinacy of germs, Journal of the Mathematical Society of Japan, 73 (2021), pp. 277–299.
  • [42] B. Osińska-Ulrych, G. Skalski, and A. Szlachcińska, Łojasiewicz inequality for a pair of semi-algebraic functions, Bulletin des Sciences Mathématiques, 166 (2021), p. 102927.
  • [43] J.-S. Pang, Error bounds in mathematical programming, Mathematical Programming, 79 (1997), pp. 299–332.
  • [44] P. A. Parrilo, Semi-definite programming relaxations for semi-algebraic problems, Math. Program., 96 (2003), pp. 293–320. Algebraic and geometric methods in discrete optimization.
  • [45] P. Solernó, Effective Łojasiewicz inequalities in semi-algebraic geometry, Applicable Algebra in Engineering, Communication and Computing, 2 (1991), pp. 1–14.
  • [46] S. Starchenko, NIP, Keisler measures and combinatorics, no. 390, 2017, pp. Exp. No. 1114, 303–334. Séminaire Bourbaki. Vol. 2015/2016. Exposés 1104–1119.
  • [47] L. van den Dries, Tame topology and o-minimal structures, vol. 248 of London Mathematical Society Lecture Note Series, Cambridge University Press, Cambridge, 1998.
  • [48] R. J. Walker, Algebraic Curves, Springer, New York, NY, USA, 1978.