• Keine Ergebnisse gefunden

Euclidean Distance Degree and Mixed Volume

N/A
N/A
Protected

Academic year: 2022

Aktie "Euclidean Distance Degree and Mixed Volume"

Copied!
23
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

https://doi.org/10.1007/s10208-021-09534-8

Euclidean Distance Degree and Mixed Volume

P. Breiding1·F. Sottile2·J. Woodcock2

Received: 22 December 2020 / Revised: 30 May 2021 / Accepted: 25 June 2021

© The Author(s) 2021

Abstract

We initiate a study of the Euclidean distance degree in the context of sparse polyno- mials. Specifically, we consider a hypersurface f =0 defined by a polynomial f that is general given its support, such that the support contains the origin. We show that the Euclidean distance degree of f =0 equals the mixed volume of the Newton poly- topes of the associated Lagrange multiplier equations. We discuss the implication of our result for computational complexity and give a formula for the Euclidean distance degree when the Newton polytope is a rectangular parallelepiped.

Keywords Euclidean distance degree·Newton polytopes·Nonlinear algebra Mathematics Subject Classification 14M25·90C26

Communicated by Teresa Krick.

P. Breiding funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Projektnummer 445466444; and funded by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (Grant Agreement No. 787840)

Research of Sottile supported by Simons Collaboration Grant for Mathematicians 636314.

B

P. Breiding

paul.breiding@mis.mpg.de F. Sottile

sottile@math.tamu.edu J. Woodcock jdubbs11@tamu.edu

1 Max-Planck-Institute for Mathematics in the Sciences Leipzig, Inselstr. 22, Leipzig 04103, Germany

2 Department of Mathematics, Texas A&M University, College Station, Texas 77843, USA

(2)

1 Introduction

LetX ⊂Rnbe a real algebraic variety. For a pointu ∈RnX, consider the following problem:

compute the critical points ofdX: X →R, xux, (1) whereu−x =

(ux)T(ux)is the Euclidean distance onRn.

Seidenberg [26] observed that ifX is nonempty, then it contains a solution to (1).

He used this observation in an algorithm for deciding if X is empty. Hauenstein [14] pointed out that solving (1) provides a point on each connected component of X. So the solutions to (1) are also useful in learning the number and position of the connected components of the variety. From the point of view of optimization, problem (1) is a relaxation of the optimization problem of finding a pointxX that minimizes the Euclidean distance tou. A prominent example of this is low-rank matrix approximation, which can be solved by computing the singular value decomposition.

In general, computing the critical points of the Euclidean distance betweenX andu is a difficult task in nonlinear algebra.

We consider problem (1) when X ⊂ Rn is areal algebraic hypersurface inRn defined by a single real polynomial,

X =VR(f):= {x∈Rn| f(x)=0}, where f(x)= f(x1, . . . ,xn)∈R[x1, . . . ,xn].

The critical points of the distance functiondXfrom (1) are calledED-critical points.

They can be found by solving the associated Lagrange multiplier equations. This is a system of polynomial equations defined as follows.

Let us writeifor the operator of partial differentiation with respect to the variable xi, so thatif := xfi, and also write∇f(x)=(∂1f(x), . . . , ∂nf(x))for the vector of partial derivatives of f (itsgradient). TheLagrange multiplier equationsare the following system ofn+1 polynomial equations in then+1 variables(λ,x1, . . . ,xn).

Lf,u(λ,x) :=

f(x)

f(x)λ(ux)

= 0, (2)

whereλis an auxiliary variable (the Lagrange multiplier).

We consider thenumber of complex solutionstoLf,u(λ,x) =0. For generalu, this number is called theEuclidean distance degree(EDD) [9] of the hypersurface

f =0:

EDD(f) := number of solutions toLf,u(λ,x)=0 inCn+1for generalu. (3) Here, “general” means for alluin the complement of a proper algebraic subvariety of Rn. In the following, when referring to EDD(f)we will simply speak of the EDD of

f.

Figure1shows the solutions toLf,u(λ,x)=0 for a biquadratic polynomial f.

(3)

Fig. 1 The curveX=

VR(x12x223x123x22+5)R2 is in blue andu=(0.025,0.2)is in green. The 12 red points are the critical points of the distance functiondX; that is, they are the x-values of the solutions to Lf,u(λ,x)=0. In this example, the Euclidean distance degree ofXis 12, so all complex solutions are in fact real (Color figure online)

Determining the Euclidean distance degree is of interest in applied algebraic geom- etry, but also in related areas, because, as we will discuss in Sect.3, our results on the EDD of f have implications for the computational complexity of solving problem (1).

There is a subtle point about EDD(f). The definition in (3) does not need us to assume thatVR(f)is a hypersurface inRn. In fact,VR(f)can even be empty. Rather, EDD(f)is a property of thecomplex hypersurface XC:=VC(f). We will therefore drop the assumption ofVR(f)being a real hypersurface in the following. Nevertheless, the reader should keep in mind that for the applications discussed at the beginning of this paper the assumption is needed. We will come back to those applications only in Sects.3.2and3.3.

In the foundational paper [9], the Euclidean distance degree of f was related to the polar classes ofXC, and there are other formulas involving characteristic classes [1] or Euler characteristic [23] ofXC. In this paper, we give a new formula for the Euclidean distance degree EDD(f).

Our main result is Theorem1in the next section. We show that, if f is sufficiently general given its supportAwith 0∈ A, then EDD(f)is equal to themixed volume of the Newton polytopes ofLf,u(λ,x). This opens new paths to compute Euclidean distance degree using tools from convex geometry. We demonstrate this in Sect.6and compute the EDD of a general hypersurface whose Newton polytope is a rectangular parallelepiped. We think it is an interesting problem to relate our mixed volume formula to other formulas involving topological invariants.

Our proof strategy relies onBernstein’s Other Theorem(Proposition1). This result gives an effective method for proving that the number of solutions to a system of polynomial equations can be expressed as a mixed volume. We hope our work sparks a new line of research that exploits this approach in other applications, not just EDD.

2 Statement of Main Results

We give a new formula for the Euclidean distance degree that takes into account the monomials in f. In Sect. 6we work this out in the special case when this Newton polytope is a rectangular parallelepiped.

Before stating our main results, we have to introduce notation: A vector a = (a1, . . . ,an)of nonnegative integers is the exponent of a monomialxa :=x1a1· · ·xnan,

(4)

and a polynomial f ∈ C[x1, . . . ,xn]is a linear combination of monomials. The set Aof exponents of monomials that appear in f is itssupport. TheNewton polytope of f is the convex hull of its support. Given polytopesQ1, . . . ,QminRm, we write MV(Q1, . . . ,Qm)for their mixed volume. This was defined by Minkowski; its defi- nition and properties are explained in [12, Sect. IV.3], and we revisit them in Sect.6.

Our main result expresses the EDD(f)in terms of mixed volume.

We denote by P,P1, . . . ,Pn the Newton polytopes of the Lagrange multiplier equationsLf,u(λ,x)from (2). That is,P is the Newton polytope of f, andPi is the Newton polytope ofifλ(uixi). Observe that P,P1, . . . ,Pnare polytopes in Rn+1, becauseLf,u(λ,x)hasn+1 variablesλ,x1, . . . ,xn.

We state our first main result. The proof is given in Sect.4.

Theorem 1 If f is a polynomial whose supportAcontains0, then EDD(f) ≤ MV(P,P1, . . . ,Pn) ,

where P is the Newton polytope of f and Piis the Newton polytope of∂if−λ(uixi) for1 ≤in. There is a dense open subset U of polynomials with supportAsuch that when fU this inequality is an equality and for u ∈Cngeneral, all solutions toLf,uoccur without multiplicity.

The important point of this theorem is that polynomial systems of the formLf,u

form a proper subvariety of the set of all polynomial systems with the same support—

its dimension is approximately 1nth of the dimension of the ambient space. We also remark that the assumption 0∈Ais essential to our proof, and it ensures thatV(f) is smooth at 0.

In the following, we refer to polynomials fUasgeneral given the supportA.

Since P,P1, . . . ,Pnare the Newton polytopes of the entries inLf,u, Bernstein’s theorem [4] implies the inequality in Theorem1(commonly known as theBKK bound;

see also [10]). Our proof of Theorem1appeals to a theorem of Bernstein which gives conditions that imply equality in the BKK bound. These conditions require thefacial systemsto be empty.

Our next main result is an application of Theorem1. We compute EDD(f)when the Newton polytope of f is the rectangular parallelepiped

B(a):= [0,a1] × · · · × [0,an], (4) wherea :=(a1, . . . ,an)is a list of positive integers. For each 1≤kn, let

ek(a) :=

1i1<···<ikn

ai1· · ·aik

be thek-th elementary symmetric polynomial innvariables evaluated ata. The next theorem is our second main result.

(5)

Theorem 2 Let a =(a1, . . . ,an). If f ∈ R[x1, . . . ,xn]has Newton polytope B(a), then

EDD(f)n k=1

k!ek(a) .

There is a dense open subset U of the space of polynomials with Newton polytope B(a)such that for fU , this inequality is an equality.

There is a conceptual change when passing from Theorem1to Theorem2. The- orem1is formulated in terms of the support of f, whereas Theorem2concerns its Newton polytope. This is because the equality in Theorem2needs the Newton poly- tope of the partial derivativeif to beB(a1, . . . ,ai−1, . . . ,an)for each 1≤in.

Whenn =2, a polynomial f with Newton polytope the 2×2 square B(2,2)is a biquadratic, and the bound of Theorem2becomes 2! ·2·2+1! ·(2+2) = 12 , which was the number of critical points found for the biquadratic curve in Fig.1.

Remark 1 Observe that for 1 ≤ i1 < · · · < ikn, if we project B(a) onto the coordinate subspace indexed by i1, . . . ,ik, we obtain B(ai1, . . . ,aik). Thus, the product ai1· · ·aik is the k-dimensional Euclidean volume of this projection and k!ai1· · ·aik is the normalized volume of this projection. On the other hand, ek(a) =

1i1<···<iknai1· · ·aik. This observation implies an appealing interpre- tation of the formula of Theorem2: It is the sum of the normalized volumes of all coordinate projections of the rectangular parallelepipedB(a).

Remark 2 [Complete Intersections] Experiments withHomotopyContinuation.

jl[7] suggest that a similar formula involving mixed volumes should hold for general complete intersections. That is, for X = {x ∈Rn | f1(x)= · · · = fk(x)=0}such that dimX = nk and f1, . . . ,fk are general given their Newton polytopes. The Lagrange multiplier Eq. (2) become f1(x)= · · · = fk(x)=0 and(ux)=0, whereλ=1, . . . , λk)is now a vector of variables, andJ =(∇f1, . . . ,fk)is the n×kJacobian matrix.

We leave this general case ofk>1 for further research.

2.1 Outline

In Sect.3, we explain implications of Theorem1 for computational complexity in the context of using thepolyhedral homotopy for solving the Lagrange multiplier equationsLf,u=0 for problem (1). In Sect.4, we explain Bernstein’s conditions and give a proof of Theorem1. The proof relies on a lemma asserting that the facial systems ofLf,uare empty. Section5is devoted to proving this lemma. The arguments that are used in this proof are explained on an example at the end of Sect.4. We conclude in Sect.6with a proof of Theorem2.

(6)

3 Implications for Computational Complexity

We discuss the implications of Theorem1for the computational complexity of com- puting critical points of the Euclidean distance (1).

3.1 Polyhedral Homotopy is Optimal for EDD

Polynomial homotopy continuation is an algorithmic framework for numerically solving polynomial equations which builds upon the following basic idea: Con- sider the system of m polynomials F(x) = (f1(x), . . . , fm(x)) = 0 in variables x = (x1, . . . ,xm). The approach to solve F(x) = 0 is to generate another sys- temG(x)(thestart system) whose zeros are known. Then,F(x)andG(x)are joined by a homotopy, which is a system H(x,t)of polynomials in m+1 variables with H(x,1) =G(x)andH(x,0)= F(x). DifferentiatingH(x,t)=0 with respect to t leads to an ordinary differential equation calledDavidenko equation. The ODE is solved by standard numerical continuation methods with initial values the zeros of G(x). This process is usually calledpath-trackingandcontinuation. For details, see [27].

One instance of this framework is thepolyhedral homotopyof Huber and Sturm- fels [16]. It provides a start systemG(x)for polynomial homotopy continuation and a homotopy H(x,t) such that the following holds: Let Q1, . . . ,Qm be the New- ton polytopes of F(x). Then, for allt(0,1]the system of polynomials H(x,t) has MV(Q1, . . . ,Qm) isolated zeros (at t = 0 this can fail, because the input F(x)=H(x,0)may have fewer than MV(Q1, . . . ,Qm)isolated zeroes). Polyhedral homotopy is implemented in many polynomial homotopy continuation software pack- ages; for instance inHomotopyContinuation.jl[7],HOM4PS[19],PHCPack [29].

Theorem1 implies that the polyhedral homotopy is optimal for computing ED- critical points in the following sense: If we assume that the continuation of zeroes has unit cost, then the complexity of solving a system of polynomial equationsF(x)=0 by polynomial homotopy continuation is determined by the number of paths that have to be tracked. This number is at least as large as the number of solutions toF(x)=0 that are computed. We say that a homotopy isoptimalif the following three properties hold: (1) the start systemG(x)has as many zeros as the inputF(x); (2) all continuation paths end in a zero ofF(x); and (3) for every zero of F(x), there is a continuation path which converges to it. In an optimal homotopy, no continuation paths have to be sorted out. That is, the number of paths which need to be tracked is optimal.

We now have the following consequence of Theorem 1, as Lf,u = 0 has MV(P,P1, . . . ,Pn)isolated solutions.

Corollary 1 If f is general given its supportAwith0 ∈ A, polyhedral homotopy is optimal for solvingLf,u=0.

Corollary 1 is an instance of a structured problem for which we have an optimal homotopy available.

(7)

In our definition of optimal homotopy, we ignored the computational complexity of path-tracking in polyhedral homotopy. We want to emphasize that this is an important part of contemporary research. We refer to Malajovich’s work [20] [21] [22].

3.2 Computing Real Points on Real Algebraic Sets

Hauenstein [14] observed that solving the Lagrange multiplier equationsLf,u = 0 gives at least one point on each connected component of the real algebraic setX = VR(f). Indeed, every real solution toLf,u =0 corresponds to a critical point of the distance function from (1). Every connected component of X contains at least one such critical point.

Corollary1shows that polyhedral homotopy provides an optimal start system for Hauenstein’s approach. Specifically, Corollary1implies that when using polyhedral homotopy in the algorithm in [14, Sect. 2.1], one does not need to distinguish between the sets E1(= continuation paths which converge to a solution toLf,u =0) and E (= continuation paths which diverge). This reduces the complexity of Hauenstein’s algorithm, who puts his work in the context of complexity in real algebraic geometry [2], [3], [24], [26].

3.3 Certification of ED-Critical Points

We considera posteriori certificationfor polynomial homotopy continuation: Zeros are certified after and not during the (inexact) numerical continuation. Implementations using exact arithmetic [15], [18] or interval arithmetic [6], [18], [25] are available. In particular, box interval arithmetic inCnis powerful in combination with our results.

We explain this.

Box interval arithmetic in the complex numbers is arithmetic with complex intervals that are of the form{x+√

−1y|x1xx2, y1yy2}forx1,x2,y1,y2∈R.

Box interval arithmetic inCnuses products of such intervals. By Theorem1, if fis gen- eral given its support andu ∈Cnis general, thenLf,uhas exactly MV(P,P1, . . . ,Pn) solutions. Therefore, if we compute MV(P,P1, . . . ,Pn)numerical approximations to solutions, and then certify that each corresponds to a true zero, and if we can certify that those true zeros are pairwise distinct, we have provably obtained all zeros ofLf,u. Furthermore, if we compute box intervals inCn+1which provably contain the zeros of Lf,u, then we can use those intervals to certify whether a zero is real (see [6][Lemma 4.8]) or whether it is not real (by checking whether the intervals intersect the real line;

this is a property of box intervals).

If it is possible to classify reality for all zeros, we can take a set of intervals {r1, . . . ,rk}ofRnwhich contain the real critical points of the distance functiondX

from (1). Therj are obtained from the coordinate projection(λ,x)xof the inter- vals containing the real zeros ofLf,u. Settingdj := {dX(s)|srj}gives a set of intervals{d1, . . . ,dk}ofR. If there existsdisuch thatdi∩dj = ∅and mindi <mindj

for alli = j, then this is a proof that the minimal value ofdX is contained indi and that the minimizer fordX is contained inri.

(8)

4 Bernstein’s Theorem

The relation between number of solutions to a polynomial system and mixed volume is given by Bernstein’s theorem [4].

Let g1, . . . ,gm ∈ C[x1, . . . ,xm] be m polynomials with Newton polytopes Q1, . . . ,Qm. Let(C×)mbe the complex torus ofm-tuples of nonzero complex num- bers and #VC×(g1, . . . ,gm)be the number of isolated solutions tog1= · · · =gm =0 in(C×)m, counted by their algebraic multiplicities. Bernstein’s theorem [4] asserts that

#VC×(g1, . . . ,gm) ≤ MV(Q1, . . . ,Qm) , (5) and the inequality becomes an equality when eachgiis general given its support. The restriction of the domain to(C×)m is because Bernstein’s theorem concerns Laurent polynomials, in which the exponents in a monomial are allowed to be negative.

An important special case of Bernstein’s theorem was proven earlier by Kush- nirenko. Suppose that the polynomialsg1, . . . ,gmall have the same Newton polytope.

This means that Q1 = · · · = Qm. We write Qfor this single polytope. Then, the mixed volume in (5) becomes MV(Q1, . . . ,Qm)=m!Vol(Q), where Vol(Q)is the m-dimensional Euclidean volume of Q. Kushnirenko’s theorem [17] states that if g1, . . . ,gmare general polynomials with Newton polytopeQ, then

#VC×(g1, . . . ,gm) = m!Vol(Q) .

That the mixed volume becomes the normalized Euclidean volume when the polytopes are equal is one of three properties which characterize mixed volume, the others being symmetry and multiadditivity. This is explained in [12][Sect. IV.3] and recalled in Sect.6.

Inequality (5) is called the BKK bound [5]. The key step in proving it is what we callBernstein’s Other Theorem. This a posteriorigives the condition under which inequality (5) is strict (equivalently, when it is an equality). We explain that.

Letg∈C[x1, . . . ,xm]be a polynomial with supportA⊂Zm, so that

g =

a∈A

caxa (ca∈C) .

Forw∈Zm, definehw(A)to be the minimum value of the linear functionxw·x on the setAand writeAwfor the subset ofAon which this minimum occurs. This is thefaceofAexposed byw. We write

gw :=

a∈Aw

caza, (6)

for the restriction ofg toAw. For w ∈ Zm and a systemG = (g1, . . . ,gm)ofm polynomials, thefacial systemisGw:=((g1)w, . . . , (gm)w).

We state Bernstein’s Other Theorem [4][Theorem B].

(9)

Proposition 1 [Bernstein’s Other Theorem] Let G = (g1, . . . ,gm)be a system of Laurent polynomials in variables x1, . . . ,xm. For each1 ≤ im, letAi be the support of giand Qi =conv(Ai)its Newton polytope. Then

#VC×(g1, . . . ,gm) < MV(Q1, . . . ,Qm)

if and only if there is0= w∈ Zmsuch that the facial system Gwhas a solution in (C×)m. Otherwise,#VC×(g1, . . . ,gm)is equal toMV(Q1, . . . ,Qm)

While this statement is similar to Bernstein’s formulation, we use its contrapositive, that the number of solutions equals the mixed volume when no facial system has a solution. We use Bernstein’s Other Theorem whenG=Lf,uandm=n+1. For this, we must show that for a general polynomial f with supportA⊂Nn, all the solutions toLf,u = 0 lie in(C×)n+1and no facial system(Lf,u)w =0 for 0 =w ∈ Zn+1 has a solution in(C×)n+1. The latter is given by the next theorem which is proved in Sect.5.

Theorem 3 Suppose that f is general given its supportA, that0∈A, and that u∈Cn is general. For any nonzerow ∈ Zn+1, the facial system(Lf,u)w has no solutions in(C×)n+1.

Using this theorem we can now prove Theorem1.

Proof (Proof of Theorem 1) Suppose that a polynomial f(x) ∈ C[x1, . . . ,xn] is general given its supportAand that 0∈A. We may also suppose thatu ∈CnVC(f) is general. By Theorem3, no facial system(Lf,u)w has a solution. By Bernstein’s Other Theorem, the Lagrange multiplier equationsLf,u=0 have MV(P,P1, . . . ,Pn) solutions in (C×)n+1. It remains to show that there are no other solutions to the Lagrange multiplier equations.

For this, we use standard dimension arguments, such as [13][Theorem 11.12], and freely invoke the generality of f. Consider theincidence variety

Sf := {(u, λ,x)∈Cnu×Cλ×Cnx |Lf,u(λ,x)=0},

which is an affine variety. As f =0 is an equation inLf,u=0, this is a subvariety of Cnu×Cλ×XC, whereXCis the complex hypersurfaceXC=VC(f).

Writeπfor the projection ofSf toXCand letxXC. The fiberπ1(x)overxis {(u, λ)∈Cnu×Cλ | ∇f(x)=λ(ux)}.

Let(u, λ)π1(f). As f is general,XCis smooth, so that∇f(x)=0 and we see thatλ=0 andu =x. Thusu =x+1λf(x). This identifies the fiberπ1(x)with C×λ, proving thatSfXCis aC×-bundle and thus is irreducible of dimensionn.

The projection ofSf toCnuis dominant, and therefore, Bertini’s theorem implies that the general fiber is zero-dimensional and smooth. That is, foru ∈ Cnugeneral, Lf,u=0 has finitely many solutions and each has multiplicity 1.

(10)

(0,1,0)

(1,1,0) (0,0,0)

(1,0,0) (2,1,0) P

(0,0,1) (1,0,1)

(0,1,0)

(1,1,0) (0,0,0)

P1

(0,0,1)

(0,1,1)

(0,0,0) (1,0,0)

(2,0,0) P2

Fig. 2 The three Newton polytopes ofLf,ufor f =c00+c10x1+c01x2+c11x1x2+c21x21x2

LetZXCbe the set of points ofXCthat do not lie in(C×)nand hence lie on some coordinate plane. As f is irreducible and f(0)=0, we see thatZhas dimensionn−2, and its inverse imageπ1(Z)inSf has dimensionn−1. The image W ofπ1(Z) under the projection toCnu consists of those pointsu ∈ Cnu which have a solution (x, λ)toLf,u(λ,x)=0 withx/(C×)n. SinceW has dimension at mostn−1, this shows that for generalu all solutions toLf,u(λ,x)=0 lie in(C×)n+1(we already showed thatλ=0).

This completes the proof of Theorem1.

4.1 Application of Bernstein’s Other Theorem

To illustrate Theorem3, let us consider two facial systems of the Lagrange multiplier equations in an example.

LetiAbe the support ofif. It depends upon the supportAof f and the indexiin the following way. Letei :=(0, . . . ,0,1,0, . . . ,0)be theith standard basis vector (1 is in positioni). To obtainiAfromA⊂Nn, first remove all pointsaAwithai =0, then shift the remaining points by−ei. The support ofi fλ(uixi)is obtained by addinge0andei+e0toiA. (As usual, we identifyNnwith{0} ×Nn ⊂Nn+1.) Throughout the paper, we associate toλtheexponent with index0.

Consider the polynomial in two variables,

f = c00+c10x1+c01x2+c11x1x2+c21x12x2.

Its support isA = {(0,0), (0,1), (1,1), (2,1), (1,0)}and its Newton polytope is P =conv(A), which is a trapezoid. Figure2shows the Newton polytope P along with the Newton polytopes of1fλ(u1x1)and2fλ(u2x2). These are polytopes inR3; we plot the exponents of the Lagrange multiplierλ in the (third) vertical direction in Fig.2.

The faces exposed byw=(0,1,0)are shown in red in Fig.3.

The corresponding facial system is

(Lf,u)w =

c00+c10x1

c10λ(u1x1) c01+c11x1+c21x12λu2

.

(11)

(0,1,0)

(1,1,0) (0,0,0)

(1,0,0) (2,1,0)

(0,0,1) (1,0,1)

(0,1,0)

(1,1,0) (0,0,0)

(0,0,1)

(0,1,1)

(0,0,0) (1,0,0)

(2,0,0) Fig. 3 The facesAw,(A1)wand(A2)wforw=(0,1,0)are shown in red (Color figure online)

(0,1,0)

(1,1,0) (0,0,0)

(1,0,0) (2,1,0)

(0,0,1) (1,0,1) (0,1,0)

(1,1,0) (0,0,0)

(0,0,1)

(0,1,1)

(0,0,0) (1,0,0)

(2,0,0) Fig. 4 The facesAw,(A1)wand(A2)wforw=(0,1,1)are shown in red (Color figure online)

Let us solve(Lf,u)w =0. We solve the first equation forx1and then substitute that into the second equation and solve it forλto obtain

x1 = −c00

c10

and λ = c10

u1x1 = c102 c10u1+c00 .

Substituting these into the third equation and clearing denominators gives the equation 0 = (c10u1+c00)(c103c11c10c00+c200c21)c410u2

which does not hold for f,u general. The proof of Theorem3is divided into three cases and one involves suchtriangular systems, which are independent of some of the variables.

The faces exposed byw=(0,−1,1)are shown in red in Fig.4.

The corresponding facial system is

(Lf,u)w =

c01x2+c11x1x2+c21x21x2

c11x2+2c21x1x2

c01+c11x1+c21x12λx2

⎦ =

fw

1(fw)

2(fw)λx2

.

Observe thathw(A)= −1 and that we have

hw(A)· fw = −fw = w1·x1·1(fw)+w2·x2·2(fw))

= 0·x1·(∂1(fw))+(−1)·x2·(∂2(fw)) = x22f. (7)

(12)

This is an instance ofEuler’s formula for quasihomogeneous polynomials(Lemma 2). If(λ,x)is a solution to(Lf,u)w=0, then the third equation becomes2f =λx2. Substituting this into (7) gives 0= −fw =λx22, which has no solutions in(C×)3. One of the cases in the proof of Theorem3exploits Euler’s formula in a similar way.

5 The Facial Systems of the Lagrange Multiplier Equations are Empty Before giving a proof of Theorem3, we present two lemmas to help understand the support of f and its interaction with derivatives of f and then make some observations about the facial system(Lf,u)w.

Let f ∈ C[x1, . . . ,xm]be a polynomial with supportA ⊂Nn, which is the set of the exponents of monomials of f. We assume that 0 ∈ A. As before we write

iA⊂Nnfor the support of the partial derivativeif. Forw∈Zn, the linear function xw·xtakes minimum values onAand oniA, which we denote by

h = hw(A) := min

a∈Aw·a and hi = hw(∂iA) := min

a∈∂iAw·a. (8) (We suppress the dependence onw.) Since 0A, we haveh≤ 0. Also, ifh=0 and if there is someaAwithai >0, thenwi ≥0.

Recall that the subsets of A andiA where the linear function xw·x is minimized are theirfacesexposed byw,

Aw := {a∈A|w·a=h} and (∂iA)w := {a ∈iA|w·a=hi}. (9) The proof below of Lemma 1 shows that i(Aw)(∂iA)w with equality when

∅ = i(Aw). As in (6) we denote by fw the restriction of f toAw, and similarly, (∂i f)wdenotes the restriction of the partial derivativeif to(∂iA)w. Theith partial derivative of fw isi(fw)and its support is denoted i(Aw). The proof below of Lemma 1 shows thati(Aw)(∂iA)wwith equality when∅ =i(Aw).

Our proof of Theorem3uses the following two results.

Lemma 1 For each1 ≤in andw∈ Zn, we have hihwi. If∂i(fw)=0, then∂i(fw)=(∂i f)w and hi =hwi.

Proof Fix 1 ≤in. LetaiA. Thena+eiAand sohw·(a+ei)= w·a+wi. Thusw·ah−wi. Taking the minimum overaiAgiveshih−wi. Suppose now that∅ = i(Aw). Letai(Aw). Thena+eiAw andh = w·(a+ei)=w·a+wi. But thenh−wi =w·ahi, which implies thathi =h−wi. It also implies thatw·a =hi. SinceAwA, we have thataiA. Asw·a =hi, we conclude thata(∂iA)w. This proves the inclusioni(Aw)(∂iA)w.

For the other inclusion, suppose thati(Aw)= ∅. As we showed,hi =hwi. Leta(∂iA)w. Thenw·a =hi and asaiA, we havea+eiA. But then w·(a+ei)=hi +wi =h, so thata+eiAw. We conclude thatai(Aw).

(13)

To complete the proof, observe thati(fw)=0 is equivalent toi(Aw)= ∅, and thati(fw)and(∂if)ware subsums ofi f over terms corresponding toi(Aw)and

to(∂iA)w, respectively.

The restriction fwof f to the face ofAexposed bywis quasihomogeneous with respect to the weightw, and thus, it satisfies a weighted version of Euler’s formula.

Lemma 2 (Euler’s formula for quasihomogeneous polynomials) Forw ∈ Zn, we have

h· fw = n i=1

wixii(fw) .

Proof For a monomialxawitha ∈Zn and 1≤in, we have thatxiixa =aixa. Thus

n i=1

wixiixa = n i=1

wiai xa = (w·a)xa.

The statement follows because foraAw(the support of fw),w·a =h. Our proof of Theorem3investigates facial systems(Lf,u)w for 0= w ∈ Zn+1 with the aim of showing that for f general given its supportA, no facial system has a solution. Recall from (2) that the Lagrange multiplier equations for the Euclidean distance problem are

Lf,u(λ,x1, . . . ,xn) =

⎢⎢

⎢⎣

f(x1, . . . ,xn)

1fλ(u1x1)

nfλ(u... nxn)

⎥⎥

⎥⎦ = 0.

Fix 0= w =(v, w1, . . . , wn)∈ Zn+1. The initial coordinate ofwisv ∈Z. It has index 0 and corresponds to the variableλ.

The first entry of the facial system(Lf,u)wis fw. The shape of the remaining entries depends onwas follows. Recall from (8) that we have seth:=min{w·a|aA}

andhi :=min{w·a|aiA}. Asvandv+wiare the weights of the monomialsλui

andλxi, respectively, there are seven possibilities for each of these remaining entries,

(∂if λ(uixi))w =

(∂if)w ifhi <min{v, v+wi}, (∂if)wλ(uixi) ifhi =vandwi=0, (∂if)wλui ifhi =vandwi>0, (∂if)w+λxi ifhi =v+wiandwi<0,

−λ(uixi) ifhi > vandwi=0,

−λui ifhi > vandwi>0, λxi ifhi > v+wiandwi<0.

(10)

Note that if one of the polynomials fw or(∂i fλ(uixi))w is a monomial, then (Lf,u)w has no solutions in(C×)n+1.

(14)

For a subsetI ⊂ {1, . . . ,n}and a vectoru ∈ Cn, letuI := {ui | iI}be the components ofu indexed byiI. We similarly writewI forw ∈ Zn andxI for variablesx∈Cnand writeCIfor the corresponding subspace ofCn.

We recall Theorem3, before we give a proof.

Theorem 4 Suppose that f is general given its supportA, that0∈A, and that u∈Rn is general. For any nonzerow∈Zn+1, the facial system(Lf,u)whas no solutions in (C×)n+1.

Proof Let 0=w=(v, w1, . . . , wn)∈Zn+1. As before,vcorresponds to the variable λandwi toxi. We argue by cases that depend uponwandA, showing that in each case, for a general polynomial f with supportA, the facial system has no solutions in(C×)n+1. Note that the last two possibilities in (10) do not occur as they give monomials. As f has supportA, if∂i(fw)=0, thenAw⊂ {a∈Nn |ai =0}.

We distinguish three cases.

Case 1 (the constant case):Suppose thati fw =0 for all 1 ≤in. Then fw is the constant term of f. Since 0∈A, this is nonvanishing for f general and the facial system(Lf,u)whas no solutions.

For the next two cases, we may assume that there is a partitionIJ = {1, . . . ,n} withI nonempty such thati fw =0 foriIandjfw=0 for jJ. By Lemma 1, we have

hi = hwifor alliI. (11) As jJ implies thatjfw =0, we see that ifaAw, thenaJ =0. This implies that fwis a polynomial in only the variablesxI, that is, fw ∈C[xI].

Case 2 (triangular systems):Suppose that foriI,wi ≥0, that is,wI ≥ 0. We claim that this implieswI=0. To see this, letaAw. As we observed,aJ =0. We have

0 ≥h = w·a = wI·aI ≥ 0.

Thush=wI·aI =0, which implies that 0∈Aw. LetiI. Sinceifw=0, there exists someaAwwithai >0. SincewI·aI=0 for allaAw, we conclude that wi =0.

LetiI. By Lemma1, we havehi =hwi, so thathi =h=0, and we also have(∂if)w =ifw. Aswi =0, the possibilities from (10) become

(∂ifλ(uixi))w =

⎧⎪

⎪⎩

ifw ifv >0,

ifwλ(uixi) ifv=0,

−λ(uixi) ifv <0.

We consider three subcases ofv <0,v >0, andv=0 in turn. Suppose first that v < 0 and that(λ,x)(C×)n+1is a solution to(Lf,u)w. Asλ = 0 and we have λ(uixi)= 0 for alliI, we conclude thatxI = uI. Since fw ∈ C[xI]is a general polynomial with supportAwandu is general, we do not have fw(uI)=0.

Thus(Lf,u)whas no solutions whenv <0.

Referenzen

ÄHNLICHE DOKUMENTE

Das Zweite ist, dass mir im Umgang mit den Schülern im Laufe meiner 20-jährigen Berufstätigkeit doch be- wusster wird, dass beispielsweise die Anzahl der Schüler, die auch

Given the fixed budget of the The National Institute for Health and Clinical NHS, the appropriate threshold is that of the oppor- Excellence (NICE) is charged with the task of

We use Erd¨ os’ probabilistic method: if one wants to prove that a structure with certain desired properties exists, one defines an appropriate probability space of structures and

The following theorem (also from Chapter 2 of slides) has an analogous formulation..

Mathematische Grundlagen der Informatik RWTH

Mathematische Grundlagen der Informatik RWTH

Let us however anticipate that, under the assumptions which will be spelled out for the existence theorem, the local slope |∂φ| will turn out to be a strong upper gradient for

Deduce from Remark 35.12 of the Lecture Notes that column orthogonality relations for the Brauer characters take the form Π tr Φ =