PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 KIProj/AccaPreleIRCIDatoDuejndd MIXED VOLUME AND TOTAL DEGREE By Ying Zhang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mathematics 2008 ABSTRACT MIXED VOLUME AND TOTAL DEGREE By Ying Zhang This thesis focuses on the study of solving several extensible benchmark polynomial systems by homotopy continuation methods. By establishing the relationship between their mixed volume and total degree, we find that for most of those systems the difference between their mixed volume and total degree is very minimal. Consequently, those systems should be solved by the classical linear homotopy method rather than the polyhedral homotopy method, although in general the polyhedral homotopy method is the typical choice for solving sparse systems. Furthermore, by restricting to the classical linear homotopy on solving those systems, we may take the special structure of the systems into account for solving the systems efficiently. This precious aspect of the classical linear homotopy does not seem to exist in the polyhedral homotopy method. To my sons, Ray and Ryan. iii ACKNOWLEDGMENTS This thesis could not be finished without the help and support of many people who are gratefully acknowledged here. First and foremost, I’m honored to express my deepest gratitude to my dedicated thesis advisor, Professor Tien-Yien Li, with whose guidance I could have worked out this thesis. He has offered me valuable ideas and suggestions with his profound knowledge in math- ematics and rich research experience. His patience and kindness are greatly appreciated. Besides, he is always willing to discuss with me anytime he is available. I have learned a lot from him not only about doing research, but also his positive attitude to life. I’m very much obliged to his efforts of helping me complete the dissertation. I’m also extremely grateful to Professor Gang Bao for his support and help when I was at Michigan State University. What’s more, I would like to thank the members of my thesis committee, Professor Andrew Christlieb, Professor Di Liu, Professor Jianliang Qian, and Professor Changyi Wang for the time they spent on serving on my committee. Thanks are also due to my fellow graduate students Tianran Chen, Tsung-Lin Lee, Rostam Sabeti, Chih-Hsiung Tsai, and Alice Wang for years of discussion and collaboration. Special thanks go to Chih-Hsiung Tsai for his time and kindness on helping me with the numerical results in this thesis. My deepest appreciation goes to my wonderful husband, Peijun Li, and my mom for their unselfish love, enduring patience, continued support and encouragement. iv TABLE OF CONTENTS LIST OF FIGURES LIST OF TABLES Introduction 1 The polyhedral homotopy 1.1 The Bernshtein theory ............................. 1.2 The polyhedral homotopy ........................... 2 Katsura—n system 2.1 Katsura-n system ................................ 2.2 Numerical results ................................ 3 Reimer-n system 3.1 Reimer-n system ................................ 3.2 Numerical results ................................ 4 Noon-n system 4.1 Noon-n system ................................. 4.2 A special start system for Noon-n system .................. 4.3 Numerical results ................................ 5 Generalized eigenvalue problem 5.1 Generalized eigenvalue problem ........................ 5.2 A special start system for generalized eigenvalue problem .......... BIBLIOGRAPHY vi vii 10 17 17 25 28 28 32 34 34 39 47 50 50 59 63 LIST OF FIGURES 5.1 The convex hull of S for n = 3. uuuuuuuuuuuuuuuuuuuuuuuu vi 2.1 3.1 4.1 4.2 5.1 LIST OF TABLES Comparison of the classical linear homotOpy with the typical start system and the polyhedral homotopy in solving Katsura-n systems. ........ Comparison of the classical linear homotopy with the typical start system and the polyhedral homotopy in solving Reimer-n systems .......... Number of linear subsystems for different grouping of constant terms. . . . Comparison of the polyhedral homotopy and the classical linear homotopy with different start systems in solving Noon-n systems. ........... The parallel speed-up for solving generalized eigenvalue problem A2: = A81: and eigenvalue problem Ax = A3: with n = 100 and n = 200. ....... vii 27 33 42 49 Introduction Polynomial systems arise very frequently in many fields of science and engineering [7], such as formula construction, geometric intersection problems, inverse kinematics, power flow problems with PQ-specified bases, computation of equilibrium states, etc. In 1977, Garcia and Zangwill [10] and Drexler [9] independently presented theorems, which suggested that homotopy continuation methods could be used to find the full set of isolated zeros of a polynomial system numerically. During the last three decades these methods have been developed into a reliable and efficient numerical algorithm for approximating all isolated zeros of polynomial systems. Let P(:r) = 0 be a system of n polynomial equations with n unknowns. Denoting P = (191, . . . ,pn) and 2: = (:51, . . . ,xn), we want to find all isolated solutions of p1(x11"‘,xn) : 0 \ pn($1,--.,xn)=0. The classical homotopy continuation method [2, 3] for solving P(:r) = 0 is to define a trivial system 62(3) 2 ((11 (11:), . . . , qn (13)) and then follow the solution curves in the real variable t from t = 0 to t = 1, which make up the solution set of H(;r, t) = (1 — t)rQ(:c) + tP(:r) = O with generic 7‘ E C \ {0}. More precisely, all the isolated solutions of P(:z:) = 0 can be found if the system Q(:r) = 0, known as the start system, is chosen properly to satisfy the following three properties: 0 Property 0. The solutions of the start system Q(:z:) = O are known; 0 Property 1. The solution set of H (2:, t) = 0 for O S t S 1 consists of a finite number of smooth paths, and each of them can be parameterized by t in [0,1); 0 Property 2. Every isolated solution of H(a:, 1) = P(a:) = O can be reached by some path originating at t = 0, that is, the path starts from a solution of the start system H(a:,0) = Q(:r) = O. A typical choice of a start system Q(a:) = 0 satisfying Property 0—2 is d gl(zla' .. awn) = (1111311 _ b1 qn(:rl, . . . ,3”) = anxg” — bn where d1, . . . ,dn are the degrees of polynomials p1(:r), . . . , pn(x), respectively, and aj, bj, j = 1, . ..,n are random complex numbers [6, 18, 23, 29, 37, 38]. The solutions of such a start system Q(:r) = 0 can be explicitly obtained and the total number of so- lutions is d = d1 x x dn, which is known as the total degree or the Bézout number of the original polynomial system P(:L') = 0 [33]. We may then find all the isolated solutions of P(:r) = 0 by following the total degree number of paths originating from solutions of the start system Q(r) = 0. But, a great majority of the polynomial systems arising in applications have fewer than, and in some cases only a small fraction of d = d1 x - - - x d, isolated zeros. We call such a system deficient. In this case, many of the d1 x x dn paths will diverge to infinity as t —> 1, and those paths become extraneous, causing highly wasteful computation. In the middle of 1990’s, a major computational breakthrough emerged in solving de- ficient polynomial systems efficiently by the homotopy continuation method. The new method, called the polyhedral homotopy method [13], takes a great advantage of the combi- natorial root count in the Bernshtein’s theory [4], which generally provides a much tighter bound, called mixed volume, for the number of isolated zeros of a polynomial system in the algebraic tori (0‘)" = (C\ {0})". When the polyhedral homotopy method is employed to solve a polynomial system, the number of homotopy paths that need to be traced agrees with twice of the mixed volume of the polynomial system. As an important consequence, when the mixed volume of a polynomial system is far less than its total degree, then solving the systems by the polyhedral homotopy method will greatly reduce the extraneous paths and thereby considerably limit the wasteful computations. However, in the core of the polyhedral homotopy continuation method, there is a some- times costly computation, namely the mixed cell computation, which provides the critically important start system that one can handle for the polyhedral homotOpy. Indeed, this mixed cell computation can become very costly for large polynomial systems. Therefore, before the polyhedral homotopy is used to solve the polynomial system at the expense of the sometimes costly mixed cell computations, a prior knowledge on the comparison of the total degree of the system and its mixed volume is highly desirable. If a substantial difference between these two numbers of the system is absent, then, of course, the system should be solved by the classical linear homotopy rather than the polyhedral homotOpy. While it has been largely admitted for grant that for most of the sparse polynomial sys- tems their mixed volume is far less than their total degree, in this thesis we analyze several extensible benchmark polynomial systems in the opposite. By establishing the relationship between their mixed volume and total degree, we find that for most of those systems the difference between their mixed volume and total degree is very minimal. Consequently, those systems should be solved by the classical linear homotopy method. Furthermore, by restricting to the classical linear homotopy on solving those systems, we may take the spe— cial structure of the systems into account for solving the systems efficiently. This precious aspect of the classical linear homotopy does not seem to exist in the polyhedral homotopy method. The thesis is organized as follows. In Chapter 1, the Bernshtein theorem is introduced along with its application on the polyhedral homotopy continuation method for solving polynomial systems, including mixed volume and mixed cell computations. In Chapter 2, 3 and 4, we study the extensible benchmark systems Katsura-n [35], Reimer—n [35] and Noon-n [31] respectively. It is shown that for each of those systems the difference between the mixed volume of the system and its total degree is very slim, if not zero. Therefore the polyhedral homotopy continuation method, widely considered the state of the art, is inappropriate for solving those systems. They should be solved by the classical linear homotopy method. Birthermore, when the classical linear homotopy is used to solve the Noon-n systems in Chapter 4, particularly illuminating is the marvelous speed-ups by choosing proper start system to recognize the symmetry structure of the system. In Chapter 5, we study the generalized eigenvalue problem An: = ABx, where A and B are n x n matrices. When this problem is considered as a polynomial system, it contains n equations in n + 1 variables. With an appended linear equation, the mixed volume of the resulting system is shown to be n, that is far less than its total degree 2". Nonetheless, from the obvious m—homogeneous structure of the system, proper start systems with n isolated solutions are always available in the classical linear homotopy. In this situation, the employment of the polyhedral homotopy method for solving this system is still unnecessary so that the mixed cell computations can be avoided. It is commonly known that very efficient algorithms for matrix eigenvalue problems, QR algorithm for Ax = Am and QZ algorithm for A2: = ABx, have been implemented and asserted in the software package LAPACK [16]. However, as the size of the matrix becomes larger, more computing resources are required. And a natural way to allocate extra computing resources efficiently is to perform independent tasks simultaneously in parallel. Since each isolated zero of a polynomial system is computed independently of all the others in the homotopy continuation method, it provides a natural environment for the parallelization. In this regard, solving very large algebraic eigenvalue problems by the homotopy continuation method in parallel offers a great perspective in contrast to highly serial QR or QZ algorithms. CHAPTER 1 The polyhedral homotopy 1.1 The Bernshtel’n theory Let P(:1:) = (p1(:r),...,pn(x)) E C[:I:] be a given polynomial system, where a: = 1 (3:1, ...,xn). Denoting :13“ = 23‘; ”.233,” with a = (a1, . . . ,an), write pl (:13) = ZaESI diam“ 191103) : ZaESn 4mm“ where 5'1, . . .,Sn are fixed subsets of N" with cardinalities kj = #S,- and 030 E C“ = C \ {O} for a E Sj,j = 1,. . . ,n. Here 53- is called the support of gay-(:13) and denoted by supp(pj). Its convex hull 623- = conv(SJ-) in R" is called the Newton polytope of pj and S = ($1,. . . ,Sn) is the support of P(2:), denoted by supp(P). For nonnegative variables A1,...,/\n and the Newton polytopes Q_,- of p, for j 2 1,. ..,n, let A1Q1+-~+ AnQn be the Minkowski sum of A1Q1, . . . , AnQn, i.e., A1Q1+'”+’\nQn={’\17'1+"'+/\n7‘n37‘jEerj=1I~wnl~ It can be shown that the n—dimensional volume, denoted Vol", of the polytope A1Q1 + + AnQn is a homogeneous polynomial of degree n in A1, . . . , An, and the coefficient of the term Al x - -- x An in this homogeneous polynomial is called the mixed volume of the polytOpes Q1, . . . , Qn, denoted M(Q1, . . . , Q"), or the mixed volume of the supports 31,. . .,Sn, denoted M(Sl, . . .,Sn). Sometimes, it is called the mixed volume of P(rr) when no ambiguities exist. The system (1.1.1) can be embedded into the system P(c, 1:) = (p1(c, 2:), . . ., pn(c, 22)), where 291(6) 113) = ZaESI cl,a$a P(c,:r) = 5 (1.1.2) 19710333) = ZaESn 92.0930 and the coefficients on with a E 5-, j = 1,. . .,n, are taken to be a set of M = k1 + H ° + kn variables. Namely, the system P(z) in (1.1.1) is considered as a system in (1.1.2) corresponding to a set of specified values of coefficients c* = (c; a) or P(az) = P(c“, :13). Lemma 1.1.1. [12] For polynomial systems P(c, :13) in (1.1.2), there exists a polynomial system C(c) 2 (91(0),. ..,gn(c)) in the variables c = (cjfl) for a E 33- and j = 1,. . .,n such that for those coefl‘icients E 2 (035,0) for which C(E) aé 0, the root count in (C*)" of the corresponding polynomial systems in (1.1.2) is a fixed number, and the root count in (C*)" of any other polynomial systems in (1.1.2) is bounded above by this number. Remark 1.1.1. Since the zeros of the polynomial system C(c) in Lemma 1.1.1 form an algebraic set with dimension smaller than M, its complement is Open and dense with full measure in CM. Therefore, with probability one, G(c*) aé 0 for randomly chosen . ,,, _ M . . . coefi‘iczents c — (cia) E C . Hence, polynomial systems P(c*,a:) in (1.1.2) with C(c") 75 0 are said to be “ in general position”. Theorem 1.1.1. (Bernshtei’n) [4] The number of isolated zeros in (C*)”, counting multi- plicities, of a polynomial system P(x) = (p1(:r), . . . ,pn(x)) with support 5 = (5'1, . . . ,5”) is bounded above by the mixed volume M(Sl, . . . , Sn). When P(z) is in general position, it has exactly M(Sl, . . . , Sn) isolated zeros in (C*)”. In [5], this root count was nicknamed as BKK bound after its inventors, Bernshtein [4], Kushnirenko [15] and Khovanskii [14]. In general, it provides a much tighter bound compared to variant Bézout bounds [30, 33]. An apparent limitation of the theorem is that it only counts the isolated zeros of polynomial system in (C*)" rather than all the isolated zeros in the affine space C”. For the purpose of finding all the isolated zeros of a polynomial system in C", a generalized version of the assertion in the theorem which counts the roots in C" is strongly desirable. This problem was first attempted in [32] by introducing the notion of the shadowed sets and a bound for the root count in C" was obtained. Later, a significantly much tighter bound was discovered in the following theorem. Theorem 1.1.2. [27] The root count in C" of a polynomial system P(x) = (p1(:r),...,pn(:r:)) with supports Sj = supp(pj), j = 1,...,n, is bounded above by the mixed volume M(31 U {0}, . . . , Sn U {0}). Corollary 1.1.1. For polynomial system P(z) 2 (pl (11:), . . . ,pn(:r)) in (1.1.1), assume all pj(a:)s have constant term, then the number of isolated zeros of P(x) in C" is bounded above by the mixed volume M(Sl, . . . ,5”) of its supports S = (5'1, . . . ,5”). When P(x) is in general position, all zeros of P(z) in C" are isolated and its total number is exactly equal to M(Sl,...,Sn). For a polynomial p(:c) = p(a:1, . . . ,xn) of degree d, denote the associated homogeneous polynomial by 31 (En ~ d a o o, — x , u I I , a 17(170, 331: mn) 0p($0 $0 The solutions of p(x) = 0 at infinity are those zeros of p in projective space I?” = {(x0,...,xn) EC"+1\(O,...,0)}/~ with x0 = 0 where the equivalent relation ~ is given by x ~ y if x = cy for some nonzero c E C. On the other hand, zeros of p(x) in C" can be identified with zeros of p in P" with x0 = 1. When the system P(x) = (p1(x),...,pn(x)) in (1.1.1) is viewed in P", namely we consider P(x0,x1, . . . ,xn) = (131 (x0,x1, . . . ,xn), . . . ,pn(x0,x1, . . .,xn)), then Theorem 1.1.3. (Bézout) If all the zeros of P(x0,x1,...,xn) in P" are isolated, then the number of those isolated zeros, counting multiplicities, equals to its total degree. Together with Corollary 1.1.1, we conclude with the following proposition. Proposition 1.1.1. For polynomial system P(x) 2 (p1 (x), . . . ,pn(x)) in general position in which all pj (x)s have constant term, assume the zeros of P(x) at infinity are all isolated, then Total degree of P(x) = Mixed volume of P(x) + number of isolated zeros of P(x) at infinity. 1.2 The polyhedral homotopy In light of Theorem 1.1.2 given in the above section, to find all isolated zeros of polynomial system P(x) = (p1(x), . . . ,pn(x)) in C” with support S = (51, . . . ,Sn), we first add the monomial x0(= 1) to those pis which do not have constant terms. Followed by choosing co- efficients of all monomials at random, a new polynomial system Q(x) = (q1 (x), . . . , qn(x)) with support 5" = (3;, . . . , 5;), where S;- = 53- U {0} for j = 1,. . . ,n, is obtained: _ -. a (11(113) — 2,65; C,a1$ _ ‘a I 0n,a1 - aESn qn(1') = Z We call such a system an augmented system of P(x). Since all those coefficients 53-,“ for a E S; and j = 1, . . . ,n are randomly chosen, this system may be regarded as a system in general position. We want to solve this system in the first place. Afterwards, this system will be used as the start system to solve P(x) = 0 via linear homotopy. Let t denote a new complex variable and consider the polynomial system Q(x,t) = (ch (x,t), . . . ,cjn(x,t)) in the n + 1 variables (x,t), where @103,” = Zaesi Elflmatwfla) (1.2.1) — xatwn(a) aESQ Cn,a 61161.30 .: Z I and the images of each 112,- : S, —> R for j = 1, . . . , n are chosen generically. For a fixed 10 to, system Q(x,t) = (d1(x,t), . . .,(jn(x,t)) can be written as . _ w (a) (Mate) = Z I(Cl. t 1 )9?“ 0651 a 0 Q(x,t0) = 5 (1.2.2) (an,at3m(a))$a- Qn($,t0) : E E a 3;, Remark 1.2.1. [17] System (1.2.2) is in general position. Now we regard Q(x, t) = 0 as a homotopy, known as the polyhedral homotopy, defined on (C"‘)" x [0, 1] with target system Q(x, 1) = Q(x). The zero set of this homotopy is made up of k homotopy paths x(1)(t), . . . ,x(kl (t). Since each rjj(x, t) has nonzero constant term for all j = 1, . . . , n, it follows from a standard application of generalized Sard’s Theorem [1] that all those homotopy paths are smooth with no bifurcations. Therefore, both Property 1 and Property 2 hold for this homotopy. However, as for Property 0, at t = 0, Q(x, 0) E 0, so those homotopy paths can not get started because the paths originating from t = 0 can not be identified. We deal with this problem with the following design. The function as = (w1,. . . ,wn) with wj : S;- —+ R,j = 1,. . . ,n, may be considered as a generic lifting on the support S, = (5/1, . . . , 8;) of Q(x), which lifts S;- to its graph Al A I - Sj = {a= (a,wj(a))|ae Sj}, J = 1,...,n. Let 0”: = (a, 1) E IR"+1 satisfy the following condition: I I I Condition A: There exists a collection of pairs {a1,a1} C 31,. . . ,{an,an} C 5;, where 11 I {a1 — all, . . .,an — an} is linearly independent and for j = 1,. . .,n (£1330 (T): (j Hé) . .. AI . I I (a, a) >(aj,oz),a E Sj \ {aj,aj}. Here (., .) stands for the usual inner product in the Euclidean space. For such d = (a, 1), where a = (0:1, . . . ,0”), let 311 =t"°1$1 yn = t—anxn- In short, we write y = t—ax with y = (y1,. . . ,yn) or x = yta. By this transformation and a: (a1,...,an) E N”, we have x“ = x‘l11 nag” =(y1t0‘1)al ---(ynt°‘n)an _ “1 an a a +~~+anan _ a (a,a) __ y] . . . yn t 1 1 _ y t . Consequently, cjj(x,t), for j = 1,. ..,n, of Q(x,t) in (1.2.1) becomes qj(yta,t) : Z 5j,a yat waa)t j(a zaez’ Cja yat a.wj(a)),(a,1)) = Z Ej,ayat<&,&> I I S. . a5 J 631- aES'.7 where ti = (a, wj(a)) and 0? = (0,1). For j: 1,...,n, let fij = min (a, Oz) was]- 12 and consider the homotopy Ha(y, t) = (h?(y,t), . . . ,hg(y, t)) = 0 on (C*)” x [0,1], where fig-”(v.0 = Pita-(war) = Z air/“5W“? I aES. J = 2 523a?!” 2 Ej.a?/“'5 H]- for a E S'; \ {aj,a;-}. Thus when t = 0 I ,5 0:5 aL1+5 ,“1:0 631 Lay 1,aly Laly hie/.0) = 2a H “(.11. 0) = I hay, 0) : ZaES; an,aya : Ema/”ya" + 5n agyan = 0 which is known as the binomial system. Proposition 1.2.1. [12] The binomial system I 51m?!“1 + 51 a! ya1 = 0. ’ l I allianyan + an (Id/”ya” = 0 9 13 has a1 -01 ka := det I (.,-..) nonsingular isolated solutions in (C*)". Proposition 1.2.2. {12/ Different a = (01,1) 6 IR"+1 that satisfy Condition A will induce difierent homotopies H “(y, t) = 0. Those diflerent homot0pies will reach different sets of isolated zeros of H 0(y, 1) = Q(y) = Q(x). Moreover, those diflerent sets of isolated zeros of Q(x) are disjoint from each other. Proposition 1.2.3. [12] The root count of Q(x) = 0 in (C*)" or the mixed volume of the augmented system of P(x) is /m_g) :ka 2: det G a I Kan—an) A key step in solving system Q(x) = 0 by the polyhedral homotopy method described above is the search of all those vectors d = (a, 1) E Rn“ as well as their associated collections of pairs COr = ({a1,a’1},. . .,{an,a:,}) which we call mixed cells with inner normal 0: that satisfy Condition A. This is one of the most time consuming parts of the polyhedral homotopy method and well developed algorithms for finding those mixed cells can be found in [11, 19, 20, 22]. After all isolated solutions of Q(x) = 0 are attained, the linear homotopy H(x, t) = (1 — t)rQ(x) + tP(x) = 0 with generic r E C* 14 will be used to solve the target system P(x) = 0 because this homotopy now satisfies all the three properties [24]. We now summarize the polyhedral homotOpy procedure for solving polynomial systems. Given polynomial system P(x) 2 (p1 (x), . . . ,pn(x)) with support 5 = (51, ...,s.,), let s’=(s[,...,s,’,) with sg=sju{0} for j=1,...,n. 0 Step 0: Initialization. Choose polynomial system Q(x) = (q1(x),...,qn(x)) with support S, = I I , . (51,...,S'n) and generically chosen coefficients cjfl, for a E S]. and J = 1,...,n. That is, {bi-73) = Z cj,axa, j=1,...,n. aeSI. J 0 Step 1: Solve Q(x) =0. 0 Step 1.1: Choose a set of real valued functions wj : S;- -» R, j = 1,...,n, their images are generic numbers. 0 Step 1.2: Find all the cells C“ = ({a1,a’1},...,{an,a:,}),j = 1,...,n I of SI = ($1,...,S;,) induced by w = (w1,...,wn) with ti = (a, 1) E 1R"+1 .. ,J .. Al I I being the inner normal of ({a1,a1},...,{an,an}) in S' = (Si,...,Sn). 0 Step 1.3: For each d = ((1,1) 6 Rn“ and its associated cell 0“ obtained in Step 1.2. 15 0 Step 1.3.1 Solve the binomial system I _ a €1,01ya1+cla’y 1 = 0 ’ 1 I Omanya" + 5 I ya" = 0 n,an in (C*)", let the solution set be X3. 0 Step 1.3.2: Let fij = (511361) forj = 1,...,n. Follow homotopy paths y(t) of the polyhedral homotopy “dial—21 06$; Cl,ay hi"(y.t) = E H“(y,t) = light, t) = Zae [1%,avatié’él’fi” S starting from the solutions in X3. Collect all the solutions of y(l) as a subset of isolated zeros of Q(x). 0 Step 2: Solve P(x) =0. Follow the homotopy paths of the linear homotopy H(x,t) = (1 -— t)rQ(x) + tP(x) = O with generic r 6 C" starting from the solutions of Q(x) = 0 obtained in Step 1 to get all the isolated solutions of P(x) =0 at t= 1. I Remark 1.2.2. To find all isolated zeros of P(x) in C", k = M(S[, . . .,Sn) homotopy paths need to be traced in both Step 1.3 and Step 2, making it 2k homotopy paths in total. 16 CHAPTER 2 Katsura—n system In this Chapter, we consider the Katsura-n system [35]. For this polynomial system, we shall prove that the mixed volume of its augmented system is equal to its total degree. Therefore, for finding all the isolated zeros of this system in C", the polyhedral homotopy method offers no advantages as regard to minimizing the number of homotopy paths one needs to trace. 2. 1 Katsura-n system Katsura—n system actually contains n + 1, rather than n, variables x1, . . . ,xn+1. It has the following forms: 17 o if n is odd, the polynomials are 2xn+1+2xn+---+2x2+x1- l 2z§+1+2xfi+m+2x§+x§—z1 2127113714.] + 2$n_1$n + ' ' ° +2$1$2 — LE2 2xn_1xn+1 + 2xn_2xn + - - - + 2x1x3 + x3 —[ x3 [ 2x2xn+1+ 2x1xn + 2x2xn_1+~-+ x H — mu; 0 if n is even, the polynomials are 2xn+1+2xn+-~+2x2+x1—1 2xi+1+2x%+---+2x§+x¥—x1 2xnxn+1+2xn_1xn +---+2x1x2 — x2 2xn_1xn+1+ 2xn_2xn + - - - + 2x1x3 + x3 -— x3 2xx +2xx +2xx_ +---+2xx —x. k 2n+l 1n 2n1 313:2 n Appending constant terms to those polynomials without them and choosing all the coefficients randomly yield the following augmented Katsura—n system 18 o if n is odd, Ci,1$n+1+ 61,2331: + - ' ' + 61.11962 + Ci,n+1131+ 01,n+2 2 .2 2 2 02,133,,“ + 02,217; + ' ' ° + C2,n$2 + 02,n+11‘1 + C2,n+2$1 + C2,n+3 C3,lznxn+1 + C3,2$n—1$n + ' ‘ ' + C3,nxl-732 + C3,n+1$2 + C3,n+2 2 C4,1$n—1$n+1 + 04,213n—2In + - ' ' + aim—1331333 + C4,n$2 + C4,n+1933 + C4,n+2 O a a 2 . Cn.1$233n+1 + Cn.21151313n + Cn.3$2$n—1 + + Cnrfibi‘axli'l + Cn’fllléxn + Gnu-{1’ o if n is even, l Ci,1$n+1+ Ci,2$n + ' ' ' + Ci,n$2 + Ci,n+1$1+ Ci,n+2 c 2 + 2 + - - . + 2 + 2 + + 2,1$n+1 C2,2-Tn Can-732 C2,n+1-731 C2,n+2$1 C2,n+3 C3,1$n$n+1 + 03,2In—11‘n + ° ' ' + 63.121311? + 03,n+1$2 + C3,n+2 C4,1$n—117n+1 + €4,217n—21‘n + ' - - + C4,n—lxlx3 + C4,n$% + C4,n+1$3 + C4,n+2 C71,133237n+1 +Gn,2$1$n+cn,3$2$n—l+”'+c n+2xn$n+2 +0 n+4$n+c n+6- "’T 2 T "~T ”'T As elaborated in the previous chapter, for the purpose of finding all isolated zeros of polynomial system P(x) in C", rather than in (C*)", the number of paths that need to be traced in the polyhedral homotopy method is twice the mixed volume of the augmented system of P(x). Therefore the major difference, in terms of the number of paths one needs to trace, in employing the classical linear homotopy or the polyhedral homotopy for solving P(x = 0 in C" lies in the comparison of the mixed volume of the augmented system of P(x) and its total degree. The following proposition shows that for a Katsura—n system these two numbers are actually the same. 19 Proposition 2.1.1. Mixed volume of augmented Katsura-n system is equal to its total degree 2". Proof. By Proposition 1.1.1, the mixed volume of an augmented polynomial system and its total degree will be the same if the polynomial system has no zeros at infinity. We will therefore prove the assertion of the proposition by showing there is no zeros at infinity for the augmented Katsura—n system by induction on n. First of all, when n = 1, the augmented Katsura—l system is reduced to 0111132 + 012301 + C13 (21 1) czlxg + (3221:? + c23x1 + 624. The zeros of this system at infinity are the zeros of its associated homogeneous polynomial C119132 + 612931 + 613130 621.13% + 6221:? + C23x1x0 + c24x3 in P2 with x0 = 0. However, solving system 01132 + 012131 = 0 621233 + 6221‘? = 0 will result in x1 = 0 and x2 = 0. Hence, system (2.1.1) has no zero at infinity. For n = 2, augmented Katsura-2 system becomes 011953 + 012132 + 013111 + C14 C21$§ + 622213 + 6231:? + 024:6] + 625 (2-1-2) 031932333 + 032161502 + 633332 + C34 20 and its associated homogeneous polynomial is 011333 + c125102 + 013931 + 0141130 2 2 2 2 62133 + 6221132 + 023.731 + C24$1$0 + 6251130 631302113 + C321511132 + 033502330 + 03435?)- Hence, zeros at infinity of system (2.1.2) is the nonzero solutions of system 61113 + 612332 + 613561 = 0 62133 + ngxg + C2333? = 0 (2-1-3) 0312:2133 + 6321:1132 = 0. If x3 = 0, system (2.1.3) becomes 0125132 + 013221 = 0 022.73% + C2322? = 0 C32x1x2 = 0 and the only solution of which is x1 = 0 and x2 = 0. When x3 aé 0, let x3 = 1 in system (2.1.3), we have W1 (1‘) = 611 + 012502 + C13131 = 1792(23) = 621 + 622% + czar? = 0 (21.4) 10203013) = 031172 + C32$1$2 = 0- It is clear that this system has no solutions because the isolated zeros of the first two equations pp1(x) = 0 and pp2(x) = 0 are all in (C*)2 and those generically chosen coefficients C31 and C32 in pp3(x) will not subject to the nonzero constraint that pp3(x) = 21 0 imposed. So the proposition is true for n = 1 and n = 2. Now suppose the augmented Katsura—(k — 1) system has no zeros at infinity. We assume I: is even. (The proof is the same for odd I: .) Then the augmented Katsura-(k — 1) system takes the form C1,1931: + €1,2Ik—1+'“+ Chic—11:2 + C1,1531 + Cl,k+1 2 + 2 + , , , + 2 2 62,193,, €2,2$k_1 Oak—11:2 + 02,1631 + 02,k+1931 + C2,k+2 C3,l$k—1$k + 03,2-Tk—217k—1 + ' ' ' + Care—11311132 + 03.16132 + 03,Ic+1 04,1xk—217k + 04,233k—317k—1 + ' ' ' + Cit—2171333 + Cit—135% + 04,1433 + C4,k+1 Ck,1$2$k + Ck,235133k—1 + Ck,3$2-’Ek—2 + ' ' ° + 0,, k+2 513%, + Ck k 4Ik—1 + 0,, k+6- ’ 2 This system has no zeros at infinity, so the system €1,193k + Cl,2$k-1+”'+ Chic—1372 + Cum c2,1xi + c2,2xi_1 + ' - - + 62,k_1113% + c2,kx¥ 03,133k—137k + 03,293Ic—217k—1 + ° ' ° + Cat—1331162 < (2.1.6) €4,1xk—233k + C4,2$k—3-’L‘k—1 + ' ' ' + Oak—2371353 + Cue—113% Ck,1$2xk + Ck,2$1$k—1 + Ck,3$2$k—2 + ' ' ° + Ck k+2$ [ .7— Mark? has no nontrivial zeros. For n = k, the zeros at infinity of the augmented Katsura—k system 22 01,1$k+1+ 61.21% + ' ° ' + 01,1632 + Ci,k+11‘1 + Cl,k+2 2 2 _ , 2 2 + 02,113,,“ + 62,2931, + ' + 62,1602 + C2,k+1$1 + C2,k+2~771 C2,k+3 C3,1$k-’Bk+1 + C3,2$k—1$k + ' ' ' + €3,k$1$2 + €3,k+1$2 + C3,k+2 2 C4,1$k—1$k+1 + €4,2xk—213k + W + C4,k—1x1$3 + C4,k$2 + C4,k+1$3 + C4,k+2 2 ck,1x3xk+1 + Ckagxgxk + Ck,3$l$k—l + ' ' ‘ + Ck k 413k ‘1' Ck k+6 $k_1 + Ck k+8 3 2 , ck+l,lx2$k+1+ Ck+122$1$k + Ck+1,3132$k_1+ . . . + C k+2$k$k+2 +6 k+4$k +6. k+6 t “11—2— 2 T k+1’T “11—2— (2.1.7) are nontrivial zeros of I C1.131%“ + 01.21% + ' ' ' + 01,k$2 + Ci,k+1$1 2 2 ... 2 + 2 C211$k+1 + C2’2xk + + €2,532 C2’k+1$1 63,1$k$k+1 + 03.212431: + ' ' ' + 03.163132 2 l €4,1xk—1$k+1 + 04212—2112 + ' ' ' + Cale—1931553 + 04,1652 (2-1-8) Ck,1$3$k+1 + Ck,2$2$k + Ck,31‘1$k—1 + ' ' ° + Ck k+4 132;), ’ 2 2 Ck+1,1$21‘k+1 + Ck+1,2$1$k + Ck+1,31‘2$k—1+“'+ CH1 k+2$k$k+2~ \ ’ 2 2 2 23 For xk+1 = O, the above system becomes 01,227), + - - - + chxg + cl,k+1x1 2 2 2 C2,233)c + ' ' ' + 62,1652 + C2,k+1$1 03212—1331: + ' ' ° + C3,k$1$2 2 l 64,2113k—2Ik + ' ' ' + C4,k—1$1x3 + 641932 Ck,2~"3233k + Ck,3$lxk—l + ' ' ' + Ck k+4$ *T Mas—N 2 Ck+1,2$1$k + Ck+1,3$2$k—1 + ' ' ' + C k+2$k$k+2 “er T (2.1.9) and the first It polynomials of system (2.1.9) have the same form as the system in (2.1.6) with randomly chosen coefficients, which has no nontrivial zeros. Consequently, system (2.1.9) itself can not have nontrivial zeros, i.e., system (2.1.7) has no zeros at infinity when 93k+1= - For xk+1 76 0, let xk+1 = 1 in system (2.1.8) and consider , 1919105): c1.1+ 61,2331: + ° ' ' + 011332 + €1,k+1$1 2 1910203) = 02,1 + 62,2110}?c + ° ' ° + 02,1633 + 02,k+1$1 PP3(IL‘) = 03,131: + €3,2xk-1xk + ' ' ° + C3,k$1$2 ( W405) = 04,1113k—1 + €4.213k—213k + ' ° ' + Oak—1551333 + 64.1652 ' 2 ppk($) = Ck,1$3 + Ckgxglk + Ck,3$133k_1 + . . . + ck k+4xk ’T 2 [ Ppk+1($) = Ck+1,1:132 + Ck+l,2xl$k + Ck+1,332xk—1+ . . . + 24 ck+1,Tk+2x k$k+2- 2T (2.1.10) First of all, no zeros x0 = (x?,xg, . . . ,xg) of the above system can have all x3 = x3 = = x2 = 0. Otherwise it would lead to a contradiction to pp1(x0) = 0 together with 0 pp2(x0) = 0. So, without loss of generality, we suppose x3 75 0. Now consider x an isolated solution of k equations PP1($1,---,Ik)=0 pp2($1,...,$k) = 0 Ppk(171.---,Ik) =0 in k variables whose coefficients are randomly chosen but then fixed (and therefore all the solutions are isolated). However, when we substitute (23?, x3, . . . ,xg) into ppk+1(x) = 0, it imposes a nonzero constraint for the coefficients Ck+1,11 . . . ’Ck+1,k 2 of ppk+1(x) since x3 95 0. This can’t occur since those coefficients are arbitrarily chosen, they do not subject to any particular constraints. Therefore system (2.1.10) has no zeros, i.e., system (2.1.7) has no zeros at infinity when xk+1 31$ 0. Thus the assertion of the proposition is valid for n = k. This completes the proof. C] 2.2 Numerical results As a comparison, we solve the Katsura—n system numerically by both the classical linear homotopy and the polyhedral homotopy, and results are listed in Table 2.1. All the com- putations here as well as in the following chapters were carried out on a Dell PC with a Pentium 4 CPU of 2.2GHz, 1GB of memory, and results presented are restricted to the systems that can be solved within 12 hours of CPU time. Recall that we use the typical 25 start system Q(x) = (q1 (x), . . . ,qn(x)) where d (11(13): 01151311" (71, with randomly chosen complex numbers a1, . . . ,an, b1, . . . , bn in the classical linear homo- topy H(x,t) = (1 — t)Q(x) + tP(x) = 0. The speed-up ratio is the ratio of the CPU time of solving the system by the polyhedral homotopy to that by the classical linear homotopy with the typical start system. Appar- ently the table shows the classical linear homotopy works much better for finding all the solutions of Katsura—n systems [21]. For instance, when n = 16 , the polyhedral homotopy takes more than 12 hours to find all isolated zeros of the system whereas the classical linear homotopy only takes 16 minutes and 25 seconds. 26 CPU time System Total Degree Speed-up ratio Linear Polyhedral Katsura— 11 2,048 113 233 2.09 Katsura— 12 4,096 263 1m22s 3.15 Katsura—13 8,192 1m06s 5m32s 5.03 Katsura— 14 16,384 2m38s 22m14s 8.44 Katsura— 15 32,768 7m03s 1h50m26s 15.66 Katsura-16 65,536 16m253 - - Katsura- 17 131 ,072 40m483 - - Katsura-18 262,144 11135m47s - - Katsura- 19 524,288 3h50m488 - - Katsura-20 1,048,576 8h58mOOS - - Table 2.1.. Comparison of the classical linear homotopy with the typical start system and the polyhedral homotopy in solving Katsura-n systems. 27 CHAPTER 3 Reimer-n system Reimer—n system [35] is a polynomial system whose mixed volume is equal to its total degree. 3. 1 Reimer-n system The general form of Reimer-n system is 22% — 21:3 + - - - + (—1)n+1223, — 1 2251* - 21:53 + - - - + (—1)"+12x§, — 1 (3.1.1) 221.",1+1 — 223+1+-~+(—1)"+12xg+1 — 1. Since polynomials in the system all have constant term, its augmented system will 28 consist of the same monomials with generically chosen coefficients: C1133? + C1233 + ' ' ' + 617,321+ Cln+1 021:1)? + 022.73% + ' ' ° + €271.73?1 + C2n+1 (3.1.2) n+1 n+1+. Proposition 3.1.1. For Reimer-n system, mixed volume = total degree = 2 x 3 x x (n+1) = (n+1)!. Proof. While exactly the same argument that was used in the proof of Proposition 2.1.1 can also be applied here, we shall provide a different proof for this proposition due to the special structure of the system. For the Reimer-n system given in (3.1.1), the supports of polynomials in the system are S1={(2,0,...,0),(0,2,...,0),...,(0,0,...,2)}={2e1,2e2,...,2en} $2 = {(3,0,...,0),(0,3,...,0),...,(0,0,...,3)} = {3e1,3e2,...,3e,,} Sn:{(n+1,0,...,0),(0,n+1,...,0),...,(0,0,...,n+1)} = {(Tl + Del? (n +1)62,. . . , (Tl +1)en}1 where for i = 1,. . . ,n, e,- = (0,. ..,0,1,0,.. . ,0) is the ith unit vector with its ith com- ponent 1 and all other components zero. Recall that for mixed cell {a1 — a’l, . . . ,an — a2} induced by the lifted support S = . . I I (51,...,Sn) with inner normal cl = (0,1) 6 Rn“, we have {a1,a1} C 31,. . . , {aman} C 29 Sn and by Proposition 1.2.3, (.,..3\ Mixed volume of the system = Z det a (an-ah) Now, for each i = 1,. ..,n, {a,-,a;} C S,- = {(i + 1)el, . . . , (i + 1)en} implies a,- - a; = (i + 1)(e[ - 822) where e] aé e’é and they are both in {e1, . . . , en}. Let K 01 — all \ { 2(6] — eé) \ (an—a2) \(n+1)(e?—e’2’)/ Then (a_2g) detA=2x-~x(n+1)det ; =2x---x(n+1)detB Kai—2} where (1 1) 61—32 \ 85’ ‘ 6'2’ I is a matrix with all arrays being either —1, 0, or 1. When det B 7A 0, then Idet B] 2 1 and consequently [det A] 2 2 x - - - X (n + 1). Thus the mixed volume of the Reimer—n system, :0 [detA], is greater than or equal to 2 x x (n +1). On the other hand, the mixed volume of any system is less than or equal to its total 30 degree. Here the total degree of the Reimer-n system is 2 X - - - x (n+1). It follows that the mixed volume of the Reimer—n system agrees with its total degree 2 x- - - x (n+ 1) = (n+1)!. Cl Corollary 3.1.1. For the Reimer—n system, there is one and only one mixed cell regardless of what sort of liftings being applied to the support (51,. . . , Sn). I I I Proof. For mixed cell {a1 — a1, . . .,an — an} of Reimer-n system, where {a1,a1} C 31,. . . , {an,a:,} C Sn, by Proposition 1.2.3, (m_g\ Mixed volume of Reimer-n system = Z det g = (n + 1)!. a \an-aizI However, from the proof in Proposition 3.1.1, for any mixed cell {a1 — all, . . . ,an — a2} (.,—.3) det g 2 (n+1)!. (1-2,) I I _ Therefore, there can be at most one mixed cell {a1 — a1, . . . ,an — an} w1th (m_g) det 3 = (n + 1)!. I \an—anI 31 3.2 Numerical results Listed in Table 3.1 is the numerical result for solving Reimer-n system by the classical linear homotopy with the typical start system Q(x) = (q) (x), . . . ,qn(x)) where d (11(33) = 313311— b], (171(3) 2 and/iii” " bn with randomly chosen complex numbers a1, . . . ,an,b1, . . . ,bn and the polyhedral homo- topy continuation method [21]. Apparently the speed-ups of the classical linear homotopy with the typical start system over the polyhedral homotopy in solving these systems shown in the table are not as dramatic as Table 2.1 shows for Katsura-n systems. A major reason is, as indicated in Corollary 3.1.1, finding only one mixed cell may not be as costly when the system is solved by the polyhedral homotopy method. 32 CPU time System Total Degree Speed-up ratio Linear Polyhedral Reimer-4 120 0.085 0.13 1.25 Reimer-5 720 0.73 0.993 1.41 Reimer—6 5,040 9.23 12.8s 1.39 Reimer-7 40,320 1m588 2m498 1.43 Reimer-8 362,880 30m43s 36m433 1.20 Reimer—9 3,628,800 7h52m4OS 8h47m42s 1.12 Table 3.1. Comparison of the classical linear homotopy with the typical start system and the polyhedral homotopy in solving Reimer-n systems. 33 CHAPTER 4 Noon-n system 4.1 Noon-n system In this Chapter, we discuss the Noon-n system [31], x1(x%+x§+-~+x,2,—1.l)+1 2 2 2 x2(x +x +---+x —1.1)+1 P(x): 1 3 n xn(x§+x§ +---+:c§,_1 — 1.1) + 1. Since polynomials in the system all have constant terms, generically choosing its coef- ficients yields the augmented system: $1(612$§ + C13333 + ' - ' + 611133121 + C10) + d1 222(C2123‘19' + 62333 + ' ' ° + anx% + 020) + d2 (4 1 1) 272%le + 6,1217% + - - - + Can—1533-1 + cno) + dn- Before relating total degree of the system to its mixed volume, we first recall certain 34 properties concerning the multiplicity of an isolated zero of a polynomial system. Simply denoting the polynomial ring C[x1, . . . ,xn] by ’P" and treating it as a vector space, we use (’P")* to represent its dual space, consisting of all the linear functionals on ’P". Definition 4.1.1. For a given polynomial ideal I C ’P" with quotient ring Pn/I, the dual space DU] of the ideal I is the set of linear functionals in the dual space (’P"/I)* with their domain extended to P". Namely, for l E (’Pn/I)* and p E ’P" l(p):=l(r) where p€r+I. An immediate consequence of this definition is, Proposition 4.1.1. For an ideal I C ’P” and l E (’P")*, .lED[1]4==>l(p)=0 VpEI. Definition 4.1.2. A subset D of the dual space (”P”)* is closed ifi leD=> l-qu quP" where linear functional 1 - q E (P")* is defined by (I'Q)p==l(qp) for p67)". Definition 4.1.3. For 3' = (j1,...,jn) 6 N8 with |j| := Zajou and for z E C”, the 35 difierential functional 0j[z] E (Pn)* with evaluation at z is defined by a [ J() 1 ( 5'” 'Z I: _ , - . J p Jl!"'3n! 8x11...6rl,” p)(2) VP E 73”- An ideal I C ’P” is called O-dimensional if all zeros of I are isolated. Definition 4.1.4. A zero 20 of a 0-dimensional ideal I C ’P" is an m-fold zero of I if there exists a closed set of m, but no more than m, linearly independent difierentiation functionals dl = 23' ajojpo] with evaluation at 20 in the dual space ’D[I] . Definition 4.1.5. For a = 1,. . . ,n the anti-difi'erentiation operators so is defined by 6;. U[z ifj >0 Saajlzli= .7 e l o 0 lfja=0 where e0 is the 0th unit vector with its oth component 1 and all other components zero and so (2 "Yjaj [20]) Z: Z 73- 8083' [20]. j j Theorem 4.1.1. [34/ In (’P")*, a subset D(zo) of differential functionals with evaluation at 20 is closed ifl dl E P(zo) => sodl E ’D(zo) for all a = 1,. . .,n. We now establish the relation between the mixed volume and the total degree of the Noon-n system. Proposition 4.1.2. For the Noon-n system, Mixed volume = Total degree —2n = 3" — 2n. 36 Proof. From Proposition 1.1.1, the above equality holds if the system in (4.1.1) has 2n zeros at infinity. Consider its associated homogeneous polynomial system 161(612233-1- 613133 + ' ' ° + 617123721 + 6101123) + (11.138 ~ 32(c21x% + 0231133 + ' ' ' + 02,1113, + 02038) + (12178 P(cco,x1, . . . ,a‘n) 2 (4.1.2) :cn(cn1a:¥+ 9.22% + - - - + cm._1a:3,_1+ 0.02:3) + dams. The zeros at infinity of system (4.1.1) are the nontrivial zeros of system (4.1.2) in I?" with 2:0 = 0, i.e., nontrivial zeros of the system 1:1(c12x3 + c13x§ + ' - - + clnwg) 332(c2lx? + c23x§ + - - - + anxfi) :vn(cn122¥ + 0,1223% + - - - + Cnn—1$%_1)- It is clear that (1,0,...,O),(0,1,0,...,0),...,(0,...,0,1) are n isolated zeros of this system. For the multiplicity of each of those solutions, we add one more polynomial 61331 4'ng2 + ' ' ' +0n33n + Cn+l to system (4.1.2), where c,-, i = 1,. ..,n + 1 are randomly chosen complex numbers, re- 37 sulting in a system of n + 1 equations in n + 1 variables 151(x0axla- ° ° axn) = 0 P(xo,:L‘1,...,.’En) : fin($0,$1,...,xn) =0 fin+l($0?x1) ' ' ' 1x71) : 0 where 151(170,$1,- - - .Ivn) = 931(612333 + 013mg + - - - + clnxg, + c10x3)+ dlxg (4.1.3) M500. 2:1. . . . mm) = :vn(cn1rc¥ + 6..sz + - - - + Gnu-13331-1 + 07.0133) + dusts?) 15n+1(330.11.-~.$n) = C1301+62302 +°"+0n$n +Cn+l' 2 1 For $0 = 0, the solutions are 212(0,_E_”c_‘1tl 0 ...,0), 22 = (0,0,—-C3é1;i,0,...,0), .., 2,, = (0,0, . . . ,0, —E"::—1). In projective space IF", these solutions are in the same equivalence class as (0,1,0,.. .,0),(0,0,1,0,...,0),...,(0,0,. ..,0,1). Let I =< P(x0,x1, . . . ,xn) > be the ideal in ’P”+1 = C[:r0,:r1, . . . ,zrn] generated by the polynomials in (4.1.3). At solution zz- = (0,...,0,——Cncz—"'Ll,0,...,0) for i = 1,...,n, we assert that the following two linearly independent differentiation functionals dz} = 000...0[Zil(:0) = P(Zz') and 0 (”1'2 = 310...0l2il(P) = 333M229 constitute a closed subset of the dual space D[I] with maximal number of differential 38 functionals with evaluation at 21-. Obviously P(zi) = 0, so by Proposition 4.1.1, (ll,1 6 D]I]. Further, dlz-2 is also in D[I] because —}5j(22') =(26j0513j230 + 3djxg)(zi) =2 0 for j = 1,. . . ,n and a - EI—Opn+l(zi) = 0- Moreover, 32 Ci00n+1 —2 ‘—.—“ t 0- 0:120 ,, 1 Pile) = 5(26101‘2' + 6di$0)(zi) = 6. NIP-d 320...0l2z'l (131') = So, by Proposition 4.1.1, 620.0(2)) ¢ D[I]. On the other hand, for j = 1,. . . ,n, - (9 - 60...010...0lzil (Pn+1) = 5mm) = Cj 74 0, .7 6 hence, aTj[z,] ¢ D[I] for all j = 1,. . . ,n. Consequently, by Theorem 4.1.1, chI-1 and (ill-2 are the only two linearly independent differential functionals with evaluation at zz- that form a closed subset of D[I]. Therefore, for each i = 1,. . . ,n, the multiplicity of 21 is two. All together, they account for 2n solutions at infinity for the system in (4.1.1). CI 4.2 A special start system for Noon—n system As Proposition 4.1.2 indicates, for the Noon-n system, the difference between the total degree and its mixed volume is 2n, which, compared with its total degree 3", becomes very slim even when n is of moderate size. Such slim difference will certainly make it difficult to enjoy the benefits that are commonly provided by using the polyhedral homotopy method 39 in solving them. Thus, the classical linear homotopy H(x, t) = (1 — t)Q(a:) + tP(a:) = 0 (4.2.4) with the typical start system Q(:r) = (q1(a:), . . . ,qn(a:)) where d (11(3) = 013111 - ()1, (171(2)) = anxg’n — bn with randomly chosen complex numbers a1, . . . ,an,b1, . . . ,bn appears to be the pr0per choice for solving those systems. On the other hand, beyond the reach of the polyhedral homotopy method, the classical linear homotopy can take a huge advantage on the strong symmetry structure existed in the Noon-n system. In the first place, the multi-homogeneous structure [36] of the system allows the choice of the start system c1(:z:1+a)(:z:2+a:3+---+:cn+fi)(x2+:z:3+---+a:n+'7) Cg($2+a)($1+$3+"'+.’L‘n+,8)(l'1+$3+'°'+$n+’7) Q($)= (4.2.5) 0n($n+01)(331+$2+"'+$n—1+fl)($1+$2+~-+xn_1+7) where c1, 02, ..., on, a, fl, and 'y are randomly chosen complex numbers. Clearly, with this choice, the symmetry in the Noon-n system is retained and with the invariance of this symmetry in the homotopy, much fewer paths need to be traced for generating the whole solution set. 40 For instance, for Noon-3 system x1(x§+ x3 — 1.1) +1 P(x) = ”(mg + 23% — 1.1) +1 3330:? + 23% - 1.1) + 1, the start system for the linear homotopy in (4.2.4) is C1031+ 00(332 + $3 + @0132 + 1‘3 + 7) QC”) = c2(:1:2 + a)($1+ 2:3 + fl)(:1:1+ 1:3 + 7) C3(133 + a)(1‘1+ x2 + mm + $2 + 7)- To attain the whole solution set of Q(:r) = 0, we only need to permute the variables :81, 322, $3 on a subset of solutions of Q(:z:) = 0. For example, the solutions of the following three linear subsystems z1+a=0 $2+$3+fi=0 $2+$3+fi=0 x1+x3+fi=0 $2+a=0 x1+$3+fi=0 $1+x2+fl=0 $1+x2+fi=0 $3+a=0 of Q(;r) = 0 are 21 = (—a,a—fi,a—fi), 232 = (a—fl, —a,a—,B), and 23 = (a—B,a—-B, —a) respectively. Obviously, solution 23 may be attained by permuting 2:1 and 2:3 in 21 or permuting 3:2 and 9:3 in 22. In fact, any one of 21,2:2, and .23 can generate all the others by permutations. This property is actually retained on the homotopy paths initiated from those three solutions of Q(:r) = 0. So we may just follow one of these homotopy paths to reach a solution of P(x) = 0 and generate the other two solutions from this solution by permutations. 41 It is clear that the above three linear subsystems of Q(a:) = 0 share one thing in common: they all have one a and two 58 as constant terms. Let us represent this set of linear subsystems of Q(:t) = 0 by (a, 25). In general, denoting (mla, mgfi, m3'y) for the set of linear subsystems of Q(:r) with m1 as, m2 fis, and m3 75 as constants terms where m1 + mg + m3 = 3, it is easy to see that solution of any one linear subsystem in the set can generate solutions of all other linear subsystems in the set by proper permutations. Among all the possible divisions of the linear subsystems of Q(:c) = 0 in this manner, there are two sets of singular linear subsystems, hence no solutions for such systems, they are (20, B) and (2a, 7). The rest of the 8 groups and the number of linear subsystems in the same group are listed in Table 4.1. constant terms number of solutions (30:) = (a, a, a) 1 (33) = ([3. 16,16) 1 (37) = (7. 7, a) 1 (a. 2/3) = (a. 3.13) 3 (a. 2'7) = (01.7.7) 3 (25. 7) = (fi. fi, '7) 3 (fl, 27) = (fl. 7, 7) 3 (a. fi. '7) = (01.47) 6 Table 4.1. Number of linear subsystems for different grouping of constant terms. Exhibited in Table 4.1, there are, in total, 21 solutions of Q(:r) = 0, and all those solutions can be generated by just solving one linear subsystem from each of those 8 groups. Note that 21 = 33 — 2 x 3, which agrees with the mixed volume of Noon-3 system given 42 in Proposition 4.1.2. For the start system Q(:r) = (q1(:c), . . . ,qn(:c)) in (4.2.5) where (11(93)=61(1‘1+a)($2+x3+~°+xn+fi)($2+$3+“-+$n+7) Q2(I)=02($2+a)($1+$3+-”+$n+fi)($1+$3+~~+$n+7) (112(3) = Onlxn + 0X11 + 1‘2 + - ' ' + 9311—1 + film + 332 + ' ' ' + xn—l + 7), let (l1(:r), . . . ,ln($)) = 0 be linear subsystems of Q(:v) = 0, where, for each i = 1,. . . ,n, l,(:c) is a linear factor of q,-(:c). We divide those linear subsystems as follows: let (mla, mgfi, m3'y) be the set of those linear subsystems in which m1, m2 and m3 of the equations in (ll(:z:),...,ln(:r)) = 0 have a, ,6 and 'y as constant terms respectively. Of m2 course, m1+m2+m3 = n. It is clear that for fixed m1, m2 and m3 there are Cg“ ~C _m1 linear subsystems of Q(:r) = 0 in (mm, mgfl, m37) and except for m1 = n —- 1, all those linear subsystems are nonsingular. To count the total number of possible different combinations of m1, m2 and m3 for which (mla, mgfl, m37) provide nonsingular subsystems, note that for fixed m1 6 {0,1,...,n — 2,n}, there are n + 1 — m1 choices for m2, and when m1 and m2 are determined, m3 = n — (m1 + m2). Therefore, in total, there are n2+3n—2 (n+1)+n+(n—1)+-~+3+1= 2 different such combinations. One of the solutions for linear subsystems (mla, mgfi, m3'y) where m1 75 n - 1 is F‘A—‘f—Jhfl (011,...,al,a2,...,ag,a3,...,a3) (4.2.6) 43 for certain a1, 02, and a3 . This solution comes from the subsystem whose first m1 equa- tions have constant term 0:, followed by the next m2 equations having constant term 3 and all the remaining equations having constant term 7. We may use this solution to generate the solution set of all linear subsystems in (mla, mgfi, mm) by permutations. For simplicity in the description, we shall replace a], a2 and 03 by 1, 2, and 3 respec- tively, so the solution 1in (4. 2. m6) becomes (1,. ,1, 2,. .,,2 3, .,3). We generate all the permutations of (1, ,1, 2,. .,,2 3,. 3) by the following steps: m1 m2+m3 m3 Step 1: Let sq1= (1, ,1, 2, ., 2) and sq2— — (2,. 2, 3,. .,.3) Step 2: For each sql and SQ2, search from left for the first smaller-larger pair and interchange them. If an interchange is made, check if there are some smaller numbers to the left of the interchange. If not, save the resulting permutation. Otherwise, move all the smaller numbers to the left of the interchange to the most left end and save the permutation. For the currently saved permutation, repeat the procedure until no smaller- larger pairs exist. Rom this step, 077,711 permutations will be produced from sql and 0,, m3 m1 permutations from SQ2. Step 3: For each permutation produced from sq], replace all the numbers labeled 2 by the numbers in the permutations from sqg and save all the possible permutations. For instance, if one of the permutations generated from sql is (1,2,1,2) and all the permutations produced by sqg are (2, 3) and (3, 2), then the resulting permutations are (1,2,1,3) and (1,3,1,2). Namely we replace 2,2 in (1,2,1,2) by (2,3) and (3,2). m1 m2 m3 A In total, there will be 0,, m1 -Cm _m1 permutations for (1,. ,1 ,2,...,2,3,...,3 . For 44 instance, when n = 4, let m1 = 2, m2 = 1, m3 = 1 for the start system (11($) = C1($1+ 01)($2 + $3 + $4 + 5)($2 + $3 + $4 + 7) (12($) = C2($2 + a)($1+ $3 + $4 + fi)($1 + $3 + $4 + 7) Q($) = Q3($) = €3($3 + a)($1+ $2 + $4 + fi)($1 + $2 + $4 + 7) (14($) = C4($4 + a)($1+ $2 + $3 + fi)($1 + $2 + $3 + 7), there are C7731 - 033ml = C2 . C21 = 12 linear subsystems (l1(r), . . . , ln(x)) = 0 of Q(r) = 0 in which (11(2),. . .,ln(:r)) = 0 have 2 a and 1 ,6 and 1 'y as constant terms. One of them is l1(:r)=a:1+a:0 L1(:1:) : l2(a:) = 2:2 + a = 0 l3(:r)=:rl+:1:2+:z:4+fi=0 14(2) =:r1 +rg+x3+7=0 and the solution for L1(x) = 0 is (—a, —a, 20 — 7, 2a — B). For simplicity, write the solution (1, 1,2,3). For the permutations of (1, 1,2,3), we follow the above steps. Let sql = (1, 1,2,2) and sqg = (2,3). All the permutations generated from sql = (1, 1,2, 2) by following Step 2 are (1,2,1,2), (2,1,1,2), (1,2,2,1), (2,1,2,1) and (2,2,1,1), and the permutation from sq2 = (2,3) is (3,2). Then by Step 3, replace all the number 23 in all the permutations generated from sql by the numbers in each of the permuta- tions produced by 8(12 and reach all the permutations for (1, 1, 2, 3). They are (1,1, 3,2), (1,2,1,3), (1,3,1,2), (2, 1,1,3), (3,1,1,2), (1,2,3,1), (1,3,2,1), (2, 1,3,1), (3, 1,2,1), (2,3,1,1) and (3,2,1,1). By replacing 1, 2 and 3 back to —a, 20 — '7, and 2a — 3 re- spectively, solutions (-a, —a, 2a — )6, 2a — '7), (—a, 2a — 'y, —a, 20 — 3), (—a, 2a — B, —a, 20: —'y),(2a -— 'y, —a, —a, 2a — 3), (2a — 5, —a, —a, 20 — '7), (—a, 20 — 'y, 20: — 45 18) -0), (—ai 2a —fi1 20 — '7: —a)1(2a —7)—a! 2a _ fit —a))(2a —:Bi _a1 20 _7) _a)1 (201—7, 2a—fl, —a, —a, ), and (Zn—)6, 201—7, —a, —a, ) are attained. They are exactly the solutions of the following remaining 11 linear subsystems in (201, Ill, 17) : L6($) 11(3) = 131+ a l2(:12) = 1132 + a l3($) = $1 +$2 +$4 +7 l4($)=$1+$2+$3+fi l1(:r) = 221+a l2(113) = $1 +$3+$4 +7 l3($) = 173 +0 l4($)=$1+$2+$3+,6 l1($) =$2+$3+$4+7 l2(.’1:)= 332 +0 l3(.’B) = 333 +0 l4($)=$1+$2+$3+fi l1(£E)=IL‘1+Oz l2(a:)=:r1+x3+a:4+7 l3($)=$1+l‘2+$4+,3 l4(:r) = 234 + a 46 L3($) = L5($) = L7($) = r-——’\———-\/———’\——‘/——’h—-—\r———’¥——-—. l1(:1:)== z1+a l2($)=$1+$3+$4 +5 [3(2) = 2:3 +0: l4(£L‘)=.’L‘1+$2+I3+’7 l1(:1:) =$2+z3+x4+fl l2(:r) = 2:2 + (1 13(2): 273 + a l4($)=$1+$2+$3+7 l1(a:)=a:1+a l2($)=$1+$3+x4+fi l3(:1.‘) =$1+$2+x4+7 14(2): :54 + Oz l1($)=$2+$3+$4+5 l2(:l:)=a:2+a l3($)=x1+xg+x4+7 l4(:z:) = .734 + a l1($)=$2+$3+$4+7 l1($)=$2+$3+$4+fi lz($)=$2+a l2($)=$1+$3+$4 +7 1110(2)) = 141(2)) : l3($)=$1+$2+$4+5 l3($)=$3+0 l4(a:) = x4 + 0: l4(:r) = $4 + a and l1($) = $2+$3+$4+7 12(w) =x1+x3 +x4 +fl 13(2) = $3 + a l4(:r) = :34 + 0:. Therefore, tracing one of the homotOpy paths of the classical linear homotopy in (4.2.4) emanated from one of the linear subsystems of 62(2)) =4 0 in (mla, mgfi, m37) is sufficient to generate all the corresponding solutions of P(x) = 0 in this group. 4.3 Numerical results Numerical results on solving Noon-n systems by different homotopies are listed in Table 4.2. Recall that the difference between the total degree and the mixed volume of Noon-n s stems is 2n. Total de ee, 3", paths need to be traced when the typical start system Y gr Q(:r) = (q1(:r), . . .,qn(:r)) where d 91(3) = “13311— b1: Q1205) = anxg" '— bn 47 with randomly chosen complex numbers a1, . . . ,an, b1, . . . , bn is used in the classical linear homotopy H(x,t) = (1 — t)Q(:I:) + tP(:r) = 0. And twice of the mixed volume, 2 x (3" — 2n) , paths need to be traced when the polyhedral homotOpy is applied. The table clearly shows that the classical linear homotopy with the typical start system works better than polyhedral homotopy because much more paths need to be followed as n gets larger for the polyhedral homotopy method. Moreover, n2+3n—2 2 traced when a. proper start system as in (4.2.5) is assigned for the classical linear homotopy. due to the symmetry of Noon-n systems, only homotopy paths need to be This number is shown in the 3rd column on the table, it is much smaller than the mixed volume. Apparently, solving Noon-n systems by choosing start system Q(:r) in (4.2.5) leads in speed by a huge margin in finding all the isolated zeros of Noon-n systems. 48 CPU time System Mixed Volume # of paths Polyhedral Typical Q(2:) Special Q(:r) noon-10 59,029 64 5m128 1m27s 1.13 noon-11 177,1 25 76 23m27s 5m32s 1.4s noon-12 531,417 89 1h28m008 27m29s 2.23 noon— 13 1,594,297 103 7h02m103 3h7m10s 3. ls noon-19 1,162,261,429 208 - - 18.53 noon-29 6.8630377E13 463 - - 2m103 noon-39 4.0525551E18 818 — - 16m488 noon-49 239299331323 1,273 - - 29m58s noon-59 1.4130386E28 1,828 — - 1h11m395 noon-69 8.3438517E32 2,483 - - 2h16m59s noon-79 4.9269609E37 3,238 - - 4h31m03s noon—89 290932121342 4,093 - - 10h17m23s Table 4.2. Comparison of the polyhedral homotopy and the classical linear homotopy with different start systems in solving Noon-n systems. 49 CHAPTER 5 Generalized eigenvalue problem By and large, polynomial system P(x) = (p1(a:), . . . , pn (1')) whose mixed volume is much less than its total degree should be solved by the polyhedral homotopy method. However, when certain m—homogeneous structure of the system is apparent, we may still choose appropriate start systems 62(2) 2 0 in the classical linear homotopy so that the number of homotopy paths that need to be traced matches the mixed volume of the system. In such situations, the polyhedral homotopy method can no longer be beneficial in reducing wasteful computations. 5.1 Generalized eigenvalue problem Consider the generalized eigenvalue problem Ax 2 A82: 50 where 011 a12 ain a21 $22 a2n A: “111 an2 arm and r 7 b11 b12 bln b21 b22 b2n B = bnl bn2 bnn are n x n matrices. This problem is actually an n polynomial equations in n+1 variables A,:I:1,...,:I:n: /\(b11$1+“-+ b1n$n)-(a11$1+-”+ aln$n) = 0, A(bn1$1+ ‘ ' ' + bnnmn)—(an1$1+ ' ' ' + 041711.11): 0. We augment the system with a linear equation c1231+~~+onzn+cn+1=0 51 where C}, . . . , Gn+1 are randomly chosen complex numbers, yielding a polynomial system of n + 1 equations in n + 1 variables Mbu$1+m+ b1n$n)-(011$1+'°°+ ain$n) (5.1.1) A(b,,1:cl + - - - + bnnxn)—(an1x1 + - - - + annxn) €1$1+"°+ 0n$n +Cn+1- Proposition 5.1.1. The mixed volume of system (5.1.1) is 71. To prove this proposition, we need to introduce the following fundamental lemmas. Lemma 5.1.1. If 20,211,. . .,2n 6 IR" are aflinely independent, i.e., 21—20, 22—20, . . . ,2,,— 20 are linearly independent, then the volume of the convex hull of those points, denoted by Voln(conv(z0, 2:1, . . . , 2n)), is (.,_,0\ 1 32—30 (.,.-..) Lemma 5.1.2. For a polynomial system P(x) = (p1(x), . . . ,pn(x)), let 51,” .,Sn be the support of p1(x), . . . ,pn(x) respectively. If 31 = 52 = = Sn = S, then M(S, . . . , S) = n! Voln(conv(S)). Proof. The mixed volume of P(x) is the coefficient of the term Al x x An in the homogeneous polynomial Voln(A1conv(Sl)+- --+/\nconv(Sn)) of degree n in A1, . . . , An. 52 Since .9le2 = :5" :5, we have Voln(A1conv(Sl) + -- - + /\nCOIlV(Sn)) =Voln(()\1 + - -- + An)conv(S')) =(A1 + - - - + An)nVoln(conv(S)) =n!Voln(conv(S))/\1 x -- - x An + other terms. Therefore, M(S, . . . , S) = n! Voln(conv(S)). E] Proof of the proposition. Consider the system W311$1+'°'+51n$n)-(&11$1+°“+ é1n$n) + 51n+1 (5.1.2) ’\(bn1$1+ ' ‘ ' + bnnmn) — (anlxl + ' ° ' + annxn) + Cinn+1 )‘(5n+llxl + ' ' ' + 5n+1n$n) + C1$1+m+ 0n$n + 0n+1 where (1,-j, bji, i = 1, . . . ,n, j = 1, . . . ,n + 1 are randomly chosen complex numbers. The supports of all the polynomials of this system are the same. Taking n = 2 for instance, the above system becomes M511$1 + l312$2) — (a11$1 + 612$2) + (”113 M521$1 + 522$2) — (521$1 + 5.2ng) + €123 (5'1'3) 3(031271 + 032222) + C1131 + 021122 + C3 with variables A, 231 , and x2. We list the points in the support in descending lexicographic 53 Figure 5.1. The convex hull of S for n = 3. order, that is 21 >181. zo for 20, 21 E Z30 if the left-most nonzero entry of the vector 21 -— zo is positive. Then S = $1 = 52 = $3 = {(0, 0, 0),(0, 0,1),(0,1,0),(1,0, 1),(1,1,0)} = {7.0, 21, 22, 23, .24}. The graph of the convex hull of S is shown in Figure 5.1. From Lemma 5.1.2, the mixed volume of system (5.1.3) is M(S, S, S) = 3!Vol3(conv(S)). To compute the volume of the convex hull of S, we divide conV3(S) into two simplices: Q1 = conv{z0,z1,22, :43}, and Q2 = conv{z0,z2,z3,z4}. Graphically it is clear from Figure 5.1, Vol3(conv(S)) = Vol3(Q1) + Vol3(Q2). For a proof of this equality, first note that Q1 and Q2 have three points 20, 2:2, z3 in common and conv{zo, 22, 23} is a face for each. A normal of the face is v =- (1, 0, —1) since (.,-1.) (.,-x.) 22—20 = 01 0 =1i+02+(—1)k- (23—20) (101) 54 Apparently, 21 and 24 lie on different sides of the face because (21,v) x (24,v) = ((0,0,1),(1,0,-1)) x ((1,1,0),(1,0,—1)) =—1x1=—1<0. Therefore, Q1 flQg = Q and conv(S) = Q1 U622, it follows that V013(conv(S)) = Vol3(Q1) + Vol3(Q2). On the other hand, 20, 21, 22, 23 are affinely independent. By Lemma 5.1.1 / ,, _ ,0 ) l 0 0 1 ) 1 1 1 V013(Q1)=§ 22—20 =§ O 1 0 :3? K 23 — 20 J \ 1 0 1 j K 22 — 20 \ K 0 1 0 \ 1 1 1 V013(Q2) = 3’, 23 — 20 = g 1 0 1 = a, K 24 — Zo ) K 1 l 0 ) thus, 2 Vol3(conv(S)) = 37. And M(S, S, S) = 3!V013(conv(S))= 3! x g = 2. For the general system in (5.1.2), the variables are A, x1,x2, . . . ,xn. Listing the points 55 in the support in descending lexicographic order yields 5=&=&=m=nfl ={(0,0,...,0),en+1,en,...,e2,61+en+1,e1+en,..,,el+62} = {20,21,22,...,22n} where 20 = (0,0, . . . ,0), and en+2_,- for i = 1, . . . ,n, 21' = e1+ez(n+1)_,- for i=n+1,...,2n. Similarly, conv(S) can be divided into n simplicies: Q1 = conv{20,21, . . . , 2n+1} Q2 = conv{20,22, . . . , 2n+2} n = COHV{ZO,Zn1"" 22nl‘ These simplicies can intersect only at their faces and for each i = 1, . . . ,n, 56 { Zz' - 20 \ K 21' \ 1 Zi+1 — 30 1 Z{+1 V01n+l(Qi) =m : = m \ Zi+n - 30 j \ Zi+n / K \ { en+2—i \ en+2—i . C _ 1 62 _ 1 2 _(n+1)! 81 +871“ _ m en+1 K en+3—i 81 + en+2—i / \ ., l l e, \ _ 1 o _ 1 _(n+1)! : _(n+1)!' Ken+1 ) Hence, Voln+1(conv(S)) = n x (n+1) (n+1)!’ and the mixed volume of system (5.1.2) is M(S, S, . . .,S) = (n +1)!Voln+1(conv(3)) ='(n +1)! x (n n Now, support S = (S1, . . . , Sn) of system (5.1.1) is a subset of the support S = (S, . . .,S) of system (5.1.2), that is, S,- Q S for all i = 1, . . . ,n. Thus 57 On the other hand, consider the system Ax — ABx = 0 (5.1.4) x1+x2+---+xn—n-—1=0 where - .- ,- -_1 211 1 100-0 211 1 121 1 020-0 121 1 A=112 1X003 0X112 1 .111”'2. L000---n_ .111"'2l and B = I , the n x n identity matrix. For these specific matrices A and B, the ze- ros of system (5.1.4) are (A, x1, x2,..., xn) = (1,2,1,1,...,1), (2,1,2,1,...,1), ..., (n,1,1,1,...,2), that is, for i = 1,...,n, the first component of the ith zero is i, the (i + 1)th component is 2 and all the others are 1. So the number of isolated zeros of system (5.1.4) in (0')" is n, and therefore M(Sl,...,§n) 2n. So, 58 5.2 A special start system for generalized eigenvalue problem The total degree of system (5.1.1) is 2", but the system has at most n isolated zeros. So the system is deficient. If the classical linear homotopy is used to solve the problem, at least 2" — n paths will be extraneous, representing huge wasteful computations. If the polyhedral homotopy is applied, 2n paths need to be followed. But when 11 becomes larger, finding all the mixed cells requires a big amount of computations . To alleviate this problem, Li et al. [17] suggested the random product homotopy and proposed a more efficient choice for the start system Q(x) = 0: Q1($) = (A + 611)($1 + 012) (12($) = (A + C21)($2 + 622) Q($) = < 5 (5.2.5) (171(37) = (A + 0n1)($n + 0112) qn+1($) = €1$1+m+ onxn + 0n+1 where qj’s i = 1,...,n,j = 1,2 and 6),, k =1,...,n+1 are randomly chosen complex numbers. It is clear that Q(x) = 0 has exactly n isolated solutions. It is proved in [17] that for this choice of Q(x) = 0, properties 0—2 hold for the linear homotopy H(x,t) = (1 — t)Q(x) + tP(x) = 0. Thus all the isolated solutions for the generalized eigenvalue problem can be found by following n paths emanating from the solutions of Q(x) = 0. 59 The generalized eigenvalue problem Ax = ABx, in particular the eigenvalue problem Ax = Ax, for B = I , has important applications in many scientific areas. It is widely known that very efficient algorithms for matrix eigenvalue problems, QR algorithm for Ax = Ax and Q2 algorithm for Ax = ABx, have been im- plemented and asserted in the software package LAPACK [16]. However, as the size of the matrix becomes larger, more computing resources are required. And a natural way to allo- cate extra computing resources efficiently is to perform independent tasks simultaneously in parallel. Since each isolated zero of a polynomial system is computed independently of all the others in the homotopy continuation method, it provides a natural environment for the parallelization. We have tested the parallel version of our homotopy algorithms on generalized eigenvalue problem Ax = ABx with augmented linear equation clxl+---+onxn+cn+1=0 where _ _ an 0.12 . . . “111 (121 022 . . . 0.2” A = an] an2 ann 60 and L . .. are randomly chosen n x n matrices and c1, . . . , cn+1 are randomly chosen complex num- bers for n = 100 and n = 200. We also tested the parallelization for eigenvalue problem, i.e. B = I, the n x n identity matrix and A is a randomly chosen matrix. Our preliminary numerical results are listed in Table 5.1. The speed-up ratio on the table represents the ratio of the CPU time of a single processor to that of multiple processors. Almost perfect speed-ups in the table illustrate the great potential in solving very large algebraic general- ized eigenvalue problems in parallel by the homotopy continuation method in contrast to highly serial QZ or QR algorithms. 61 CPU time Speed-up ratio n # of processors Ax=ABx Ax=Ax Ax=ABx Ax=Ax 100 1 86.5 54.6 1 1 2 43.7 27.5 1.99 1.99 3 28.9 18.2 2.99 3 5 17.9 11.2 4.83 4.88 7 12.6 7.9 6.86 6.91 200 1 1795 1146 1 1 2 898 573 2 2 3 598 384 3 2.98 5 366 233 4.9 4.92 7 260 166 6.9 6.9 Table 5.1. The parallel speed—up for solving generalized eigenvalue problem Ax = ABx and eigenvalue problem Ax = Ax with n = 100 and n = 200. 62 BIBLIOGRAPHY [1] R. Abraham and J. Robbin: Transversal Mappings and Flows. W. A. Benjamin, Inc., New York-Amsterdam, 1967. [2] E. L. Allgower and K. Georg, Numerical Continuation Methods, an Introduction, Springer Series in Comput. Math, Springer-Velag, Berlin, Heidelberg, 13 (1990). [3] E. L. Allgower and K. Georg, Continuation and path following, in: Acta Numerica. 1993, A. Iserles, ed., Cambridge Univ. Press, Cambridge, 1993, 1—64. [4] D. N. Bernshtein, The number of the roots of a system of equations, Funct. Anal. Appl., 9 (1975), 183—185. [5] J. Canny and J. M. Rojas, An optimal condition for determining the exact number of roots of a polynomial system, in: Proceedings of the 1991 International Symposium on Symbolic and Algebraic Computation, ACM, 1991, 96—101. [6] S. N. Chow, J. Mallet-Paret and J. A. Yorke, Homotopy method for locating all zeros of a system of polynomials, in: Functional Differential Equations and Approximation of Fixed Points, H. O. Peitgen and H. O. Walther, eds., Lecture Notes in Mathematics, Springer-Velag, Berlin, Heidelberg, 730 (1979), 77—88. [7] H. H. Chung, polynomial homotopy and its applications to polynomial system solving, PH.D. dissertation, Michigan State University, 1998. [8] D. Cox, J. Little and D. O’Shea, Ideals, Varieties and Algorithms, An Introduction to Computational Algebraic Geometry and Commutative Algebra, Springer-Verlag, New York, 1998. [9] F. J. Drexler, Eine methode zur Berechnung siimtlicher Lbsungen von Polynomgle— ichungssystemen, Numer. Math, 29 (1977), 45-58. [10] C. B. Garcia and W. I. Zangwill, Finding all solutions to polynomial systems and other systems of equations, Math Programming, 16 (1979), 159—176. [11] T. Gao, T. Y. Li and M. Wu, MixedVol: A software package for mixed Volume com- putation, ACM Transactions on Math. Software, Vol. 31, No. 4 (2005), 555-560. 63 [12] B. Huber and B. Sturmfels, A polyhedral method for solving sparse polynomial sys- tems. Math. Comp, 64 (1995), 1541—1555. [13] B. Huber and B. Sturmfels, Bernshtein’s theorem in affine space, Discrete Comput. Geom., 7 (1997), 137-141. [14] A. G. Khovanskii, Newton polyhedra and the genus of complete intersections, Funct. Anal. Appl., 12 (1978), 38—46. [15] A. G. Kushnirenko, Newton polytopes and the Bézout theorem, Einct. Anal. Appl. 10 (1976), 233—235. [16] http://www.netlib.org/lapack/ [17] T. Y. Li, Solving polynomial systems by the homotopy continuation method, Hand- book of numerical analysis, XI (2003), 209—304, [18] T. Y. Li, On Chow, Mallet-Paret and Yorke homotopy for solving systems of polyno- mials, Bull. Inst. Math. Acad. Sinica 11 (1983), 433—437. [19] T. Y. Li and T. Gao, Mixed volume computation via linear programming, Taiwanese J. of Math. 4 (2000), 599—619. [20] T. Y. Li and T. Lee, Mixed volume computation, A revisit, Submitted. [21] T. Y. Li, T. Lee and C. Tsai, HOM4PS—2.0: A software package for solving polynomial systems by the polyhedral homotopy continuation method, Submitted. [22] T. Y. Li and X. Li, Finding mixed cells in the mixed volume computation, Found. Comput. Math. 1 (2001), 161-181. [23] T. Y. Li and T. Sauer, A simple homotopy for solving deficient polynomial systems, Japan J. Indust. Appl. Math. 6 (1989), 409—419. [24] T. Y. Li, T. Sauer and J. A. Yorke , The cheaters homotopy: an efficient procedure for solving systems of polynomial equations, SIAM J. Numer. Anal., 26 (1989), 1241-1251. [25] T. Y. Li and C. Tsai, HOM4PS—2.0para: Parallelization of HOM4PS—2.0 for solving polynomial systems, Submitted. [26] T. Y. Li and X. Wang, Nonlinear homotopies for solving deficient polynomial systems with parameters, SIAM J. Numer. Anal. 29 (1992), 1104—1118 [27] T. Y. Li and X. Wang, The BKK root count in C", Math. Comp. 65 (1996), 1477-1484. [28] P.Meravy, Symmetric homotopies for solving systems of polynomial equations, Math. Slocava, 39 (1989), 277—288. 64 [29] A. P. Morgan, A homotopy for solving polynomial systems, Appl. Math. Comput. 18 (1986), 173—177. [30] A. P. Morgan and A. J. Sommese, A homotopy for solving general polynomial systems that respect m—homogeneous structures, Appl. Math. Comput. 24 (1987), 101—-113. [31] V. W. Noonburg, A neural network modeled by an adaptive Lotka—Volterra system, SIAM J. Appl. Math. 49 (1989), 1779—1792. [32] J. M. Rojas, A convex geometric approach to counting the roots of a polynomial system, Theoret. Comput. Sci. 133 (1994), 105—140. [33] I. R. Shafarevich, Basic Algebraic Geometry, Springer-Verlag, New York, 1977. [34] H. J. Stetter, Numerical Polynomial Algebra, SIAM, 2004 [35] C. Traverso, The PoSSo test examples, available at: www.inria.fr/saga/ POL. [36] J. Verschelde and A. Haegemans, The GBQ—Algorithm for Constructing Start Systems of Homotopies for Polynomial Systems, SIAM Journal on Numerical Analysis, Vol. 30, No. 2 (1993), 583—594 [37] A. H. Wright, Finding all solutions to a system of polynomial equations, Math. Comp. 44 (1985), 125—133. [38] W. Zulener, A simple homotopy method for determining all isolated solutions to poly- nomial systems, Math. Comp. 50 (1988), 167—177. 65