'1‘!" "3"“ ”Into!” 1",. H?" L 1.175.111 J “'41.”: \‘ '-‘: "Pal-"3” "at 3'1'I1I 1‘3"“. ;"""""‘A 39111 "W . 5" 1 v 1'1 "‘ "as“ .1 r"31" I. ll H'H'Inr £1?" ”*5";\L:‘V 1'3}- , 1 :‘f 3’ . 3}." v ‘. H 3' [Wu l 1 a1 ‘5'! 1_ I ._ "MI ". K "“37",!“ “If: 11 "1-3": 'g“:1"fl$\: fro.” "3-, :'fl:"fi"l 19"." 11‘" '15-‘31‘6 ”‘3' ' ”pr‘fl ("“1“ "int“ 1,“ '2 "‘11V;'-"". ' In}; "21“"- u r" "' ""23"": Dill-"4"";j‘i1—rvpflaw-‘.'."}I"‘"-""‘"3‘41' 17'1""‘juanAV‘KrV' 01'1” 1. L~15~11."1'1 1111.111, 11:», ":33 )p‘é ¥"""I “I I I ""31"“ IQ”; E"I;\"3\33.T'1" :IliIfllI' ""1"“ $1312}, 3‘; ‘}r:‘ a- ,$ “"7” 0'1 (""5" “11““: NM 912331-11»? '0} 33.3151; "”3 " A“ 1 Ema 1 11 ‘l tn'I 1W‘3V'H 45:“... 3""‘3L'H'1I. " “""“' '1‘? 111‘ Jr 1 1 . - .1.'J'\2'1E.§§‘,'11'1'4.1111""'1'"‘1{(I '3'91'1">"1".;'d'ai, 1 u . 1:11. t- a I "1""; "7‘03““ p. :l‘",""1'"¢"f"lu >N".'.'_"fi;fi V 1:3 r0133" "‘3" l‘ TV ' 3‘1“: a?" 1." I .71! 1 “K1 9' ‘ ILL." I‘I :"'j}$."q:" ‘1‘: ”E: "1«I\' .M ”fffitiuflu‘ If v‘xI: "133 I :.:.'..:' 1‘ ‘35:“! -L .1 '1 .31 _ .1: I9; ‘ )5: 1:1" "3).: _.AL'\' u ""3"!“ ‘1...” V: 3' "1% . 1113'“. 1I ivy-1L . Lf i'":"II I |.. x V‘ '1'.) f‘“:“ “v.1!- 1 1 1' 1 _ ' “1. '. "11 1 ' '1' I ‘ _ n I 9 5 1371' 1:.lw" 1' 1 n , ,1 1| 1 1 1 . 1111 31" 1.21111 "131.13 3:" 2‘ V1. 11',II\AIH ‘1) (311-1 ll' iii-L1 "I "u“. all i *2 LI}. " EJCINA I ., f . "‘ 3‘ h' Y ‘ -. . meCIlirl’r' . . ’ £1“ WWW?“- " 'WUM'h.‘ THESIS This is to certify that the thesis entitled Existence and Computation of Static Equilibria in Certain Economic Models with Application to the PIES Model of the Energy Sector presented by Paul Arthur Rubin has been accepted towards fulfillment of the requirements for Ph. D. degree in Mathematics MP. [W Major professor Date 5 980 0-7639 'ovenou rut: : ' 25¢ per day por tu- - RETHRNING LIBRARY MATERIALS: Phco in book return to move charge from circulation records EXISTENCE AND COMPUTATION OF STATIC EQUILIBRIA IN CERTAIN ECONOMIC MODELS, WITH APPLICATION TO THE PIES MODEL OF THE ENERGY SECTOR BY Paul Arthur Rubin A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mathematics 1980 J; / afloV ABSTRACT EXISTENCE AND COMPUTATION OF STATIC EQUILIBRIA IN CERTAIN ECONOMIC MODELS, WITH APPLICATION TO THE PIES MODEL OF THE ENERGY SECTOR BY Paul Arthur Rubin In this thesis we examine the Project Independence Evaluation System (PIES) integrating model, a static equilibrium model of the energy sector of the national economy. After carefully formulating the equilibrium problem, we establish conditions sufficient to ensure the existence of an equilibrium, as well as conditions sufficient to guarantee uniqueness of that equilibrium. We study the PIES algorithm, prove that in certain cases it must converge to an equilibrium, and exhibit an example on which it fails to converge. The same is done for the PIES—VAR variant algorithm. We pose a minimization problem related to the task of locating equilibria, and propose a subgradient- based algorithm for that problem. Finally, we describe the implementation of the subgradient algorithm and discuss the results of some computational trials. To all who waited. ii ACKNOWLEDGMENTS I would like to thank my advisor, Professor V. P. Sreedharan, for his guidance in the research leading to this thesis. I am also grateful to the members of my guidance committee for their assistance, and in particular to Professor P. K. Wong for his careful reading of the dissertation. I further thank the faculty and staff of the Department of Mathematics at Michigan State University for their support of my efforts as a student. Finally, I wish to express my gratitude to Mrs. Mary Reynolds and Ms. Cindy Balzer for their excellent work in preparing the manuscript. iii List of Chapter Chapter Chapter Chapter Chapter Chapter Chapter Chapter List of TABLE OF CONTENTS Tables 0 Introduction I The Integrating Model II Existence of Equilibrium III Uhiquness of Equilibria IV The PIES Algorithm V The PIES-VAR Algorithm VI A Subgradient Projection Algorithm VII Implementation of the Subgradient Projection Algorithm References iv page page page Page page Page Page Page Page Page 21 33 42 65 76 101 110 LIST OF TABLES Table 1 Page 51 CHAPTER O INTRODUCTION The Project Independence Evaluation System [5 ] henceforth denoted PIES, is an aggregation of models Which describes the energy sector of the national economy, developed by the Federal Energy Administration (now part of the Department of Energy) as a tool for the evaluation of policy decisions. The various components of PIES model the production, refinement, conversion, transportation and consumption of a variety of energy commodities. Of particular importance is the estimation of a static partial equilibrium for the energy sector, a vector of prices at which supply and demand will be in agreement. The component models take as parameters factors such as tax policies and the pricing of crude oil by foreign producers. The static equilibrium predicted by PIES based on specified values for these parameters is taken as an indication of the expected market response to those policies. Central to the estimation of the static equilibrium is the integrating model [ 5,I7], which computes the 2 equilibrium based on supply and demand functions generated by other component models. The integrating model employs an iterative technique which successively refines approxi- mations to the equilibrium demand vector. Although observed to converge rapidly in the examples reported by Hogan and Wagner, the algorithm has evaded to date a complete theoretical analysis regarding convergence, al— though it has motivated a significant amount of research and several variant algorithms [3 ,6 ]. In this paper we attempt to shed some light on the existence and deter- mination of equilibria. In particular, we propose a new algorithm for computing the equilibria. We begin with a mathematical formulation of the integrating model, introducing the necessary concepts from convex analysis. We then prove the existence of an equilib— rium under fairly general conditions, conditions consistent with the PIES models. We note that the authors of PIES, in their publications, have assumed rather than demonstrated the existence of equilibria. uniqueness of the equilibrium has been demonstrated [15] under a number of hypotheses made by the PIES modellers. We show that one of these hypotheses appears to be inconsistent with the form of the demand model. and that in the absence of that hypothesis the equilibrium need not be unique. We examine the PIES algorithm for computing the equilibrium, exhibiting an example in which the algorithm fails to converge to a solution. This explains the failure of others to produce general convergence results, and suggests the need for either additional hypotheses or a different algorithm. We also examine one variant of the PIES algorithm, the PIES-VAR algorithm [6»], and eXhibit an example in which it fails to converge. We suggest an algorithm for solving the equilibrium problem in essentially the same form as that assumed by the PIES algorithm. The algorithm we present is applicable to a general class of problems. The proof of convergence of the algorithm Which we present assumes that the demand function has a potential. Implementation of the algorithm does not require the existence of such a potential, which suggests the distinct possibility of proving that the algorithm converges even where the demand function has no potential. We shall return to this problem at a later date. CHAPTER I THE INTEGRATING MODEL We consider the problem of determining a static market equilibrium in a context slightly more general than that of the PIES algorithm [ 5]. Let us assume that we have a finite collection of goods indexed by the integers l,...,d. In the PIES models, goods repre- sent various energy products, differentiated by type of energy, region of production and region of consumption. Wagner [17] suggests that d = 54 is typical for the PIES model. We represent demands, supplies and prices of these goods as vectors of length d, i.e., elements of the d-dimensional euclidean space I51. Let us pause to establish some notation. We do not, in general, distinguish between row and column vectors: the context determines the shape of the vector. In particular, the usual inner- product of two vectors will be denoted by the juxtaposition of those two vectors; that is, for x,y 6 Rd , 2 xy = '24 xiyi. l=l We denote by R3 the nonnegative orthant [x e 151: xi 2,0, 1 = 1,...,d}. When ordering vectors x,y E Eta, we adopt the following notation: X2y iff Xi2yi, i=l,...,d; x > y iff x 2.y and x #’y; loooopdo x >> y lff Xi > yi. 1 For any subset S of I51, we denote by int S and bd S the interior and boundary respectively of S. The affine d subspace of II of least dimension containing 8 is the affine hull of S, denoted aff S. The interior and boundary of S in the subspace topology of aff S are the relative interior and relative boundary of S, de— noted rel int S and rel bd S respectively. As much of our interest will focus on sequences of vectors rather than components of those vectors, we for convenience index members of a sequence of vectors with subscripts. Thus the symbol xn may represent the nth component of a vector x or the nth member of a sequence of vectors, depending on context. When such usage is potentially ambiguous, we will exercise greater care. To continue with our model, we assume that we are given two vector-valued functions of a vector variable, p,p'-1 : int IRS -* int IRE , which are C1 and are inverses of each other. The functions 6 p and p-1 represent respectively the vector of prices at which a specified vector of quantities is demanded (the indirect demand function) and the vector of quantities demanded at a given vector of price levels (the direct demand function). In the PIES models these functions are predicted by econometric methods and the predictions then approximated: the actual demand model is time-dependent, whereas the input fed to the integrating model represents a cross-section of the demand function with time fixed. The form chosen by the PIES modellers is the log-linear form, expressed by the equations d m.. p. = k. H q.13 , i = 1,...,d (1.0.1) 1 1 3:1 3 where pi and qi are respectively the price of and demand for the ith commodity. Adopting the notation d log x = (log x1.....log xd) for x 6 int Ig_, we can rewrite equations (1.0.1) as log p = Ki-M log q (1.0.2) where K = (log k1,...,log kd) and M is the d) -0. for all z 6 mm. Since we have assumed that f(b) is finite, by the strong duality theorem of linear programming the dual problem (1.5.2) has a maximal solution, and any maximal solution § of (1.5.2) satisfies f(b) = 5113. (1.5.3) We now show that every maximal solution y of (1.5.2) belongs to af(b). Observe that for any 2, § is feasible in the problem l3 yA g_c, y‘Z O, yz(max) (1.5.4) dual to (1.5.1). By the weak duality theorem, we have that m f(z) 2,§z for all z e Ii . (1.5.5) In view of (1.5.3) and (1.5.5), f(z) 2 f(b)+y(z-b) for all z e Rm, (1.5.6) showing that y E Bf(b). We now show that if § 6 Bf(b) then § is a maximal solution of (1.5.2). We are given that (1.5.6) holds; from (ii), it follows that for z g.b, f(b) 2 f(z) 2 f(b)+y(z-b), and so y(z-—b) g_0 for all z g.b. This shows that § 2 0. Taking z = Ax (x.2 0) in (1.5.6), we have f(Ax)-f(b ‘2 §(Ax-b). Also, from (1.5.1) f(Ax) g cx, and so cx-f(b) 2_y(Ax-b) for all x‘2 0. Since f(b) is finite, there exists a minimal solution 2 of (1.5.1) when 2 = b, with f(b) = oi. Thus cx-cx 2_y(Ax-b) for all x‘z O, 14 and so (c-§A)x 2_c§-§b for all x 2.0. (1.5.7) If the jth component of c-yA were negative, we could violate (1.5.7) by taking x to be a sufficiently large multiple of the jth standard basis vector in If}. Hence (1.5.7) implies that c-yA‘2 0, which, together with the observation that § 2_O made above, shows that i is feasible in (1.5.2). Taking x = o in (1.5.7), we see that Ci g_yb. The opposite inequality also holds, by the weak duality theorem, and so ex = yb, indicating that § is optimal in (1.5.2). 1.6 Theorem. Let A be an m> —w for all z. This means 9 is finite on eff dom g, and so it suffices to find h polyhedrally convex such that g(z) = h(z) whenever g(z) is finite. Let F be the set of feasible solutions to (1.8.4), i.e., F = [(y.S) =Y 2,0: 5 2.0: ij'SB.S C}: Which we have shown to be nonempty. Let 2 be any point Where 9 is finite. Since F is polyhedrally convex and (by virtue of the nonnegativity constraints) line-free, problem (1.8.4) has Optimal solutions, at least one of which must be an extreme point of F. F has finitely many extreme points, which we may enumerate as {(ui,vi): i=1,...,T]. Let h (w) = u.w4-v.b, i = 1,...,T, w 6 IE“, 1 1 l and let h(w) = max{hi(w) :i = 1.....T), w 6 mm. Since (1.8.4) is Optimized at one Of the extreme points Of F, by the strong duality theorem g(z) = max{uiz+vib : i = l,...,T} = h(z) whenever g(z) is finite, which is the desired result. 19 We remark that equality (1.8.2) does not extend beyond eff dom 9. Also, in the course of the proof we showed 9 to be prOper if it is ever finite. 1.9 Corollagy. The restriction of g to eff dom g is continuous. Proof: This follows from (1.8.2) and the observation that h is continuous on all Of If“. 1.10 Proposition. Let g be as in Theorem 1.8, with G = eff dom g. The set G is polyhedrally convex and closed. Proof: Let X: {xean:Bx2b,x20}. X is polyhedrally convex and G=A(X)‘]Rin I so G is polyhedrally convex, and consequently closed. We have now Obtained the desired description of Bv(g), v the supply cost function defined by (1.0.3). Let g 6 eff dom v. Replacing a,B,b and g in Theorem 1.6 with q,-B,-b and v respectively, we find that Bv(q) is nonempty and u E av(q) iff there exists 5 E If: such that uqz 0, 5‘2 0, uA-sB g_c, uq-—sb = v(q), i.e., u E av(q) iff (u,s) is an Optimal solution to the problem dual to (1.0.3) for some 8. Also, by 20 Theorem 1.8, v is polyhedrally convex on eff dom v, which by PrOposition 1.10 is polyhedrally convex and closed. We thus have a procedure for computing the subgradients of v. The time has come to face the issue of free disposal of excess goods. In view of definition 1.4 and the hypothesis that the indirect demand function p maps int Rf into int I§?, we must have some reasonable hOpe that av(q) contains at least one nonnegative vector for many (preferably all) q in eff dom v. In our formulation, Theorem 1.6 guarantees that d av(q) C Ig.. In the absence of free disposal, however, the output con- straints Of (1.0.3) become equalities. The corresponding change in Theorem 1.6 would be to write Ax = z in (1.6.1) and drOp the restriction Y.2 O in (1.6.2). In this event, we no longer have restrictions on the signs of the marginal prices. Economic arguments can be made for the nonnegativity of the marginal costs, but close scrutiny shows that such arguments require free disposal of excesses if they are to be valid. We thus assume free disposal in order to ensure nonnegative prices. In the next chapter, we exhibit conditions under which an equilibrium exists. CHAPTER II EXISTENCE OF EQ UI LI BRI UM We turn to the task of establishing the existence of an equilibrium demand, i.e., a vector q satisfying under conditions sufficiently broad to include the supply and demand models in PIES. We accomplish this in two stages. Recall that the log-linear indirect demand func- tion p of the PIES model, though continuous on int IRE , is undefined on bd IRE . We begin with the case in which p is continuous on all of RE . Our original statement of the first theorem used the conjugate function for v [11] to arrive at a formu- lation under which the Kakutani fixed-point theorem could be invoked. We present instead a version which combines a more general statement with a more direct proof. 2.1 Lemma. Let f be a convex function on IRd; then the set of minimizers of f is a convex (possibly empty) subset of Ed . 21 22 Proof: Let _ - d m — 1nf{f(x) :x 6 1R 3. The set Of minimizers of f is precisely the set d {x 6 1R : f(x) gm}. Due to the convexity of f this set is convex. 2.2 Theorem. Let K C 351 be compact, convex and nonempty, f : K 4 Rd continuous and g : K 4 IR con- tinuous and convex. Define g(x) = +w for x E K. Then there exists a point a 6 K such that f(a) E Bg(a). Proof: For each k E K, define fk : K 4 JR by fk(x) = g(x)-f(k)x for all x E K. Each f is continuous and convex on K. Since K is k compact, the set F(k) Of minimizers of fk over K is nonempty; that is, if = min f (x) mk xEK k then F(k) = film) a! 91. In view of Lemma 2.1, F(k) is also convex. Since fk is continuous, F(k) must be closed. Thus for each k 6 K, F(k) is a nonempty, compact, convex subset of K. We now show that the point-tO-set map F satisfies the hypotheses of Kakutani's fixed-point theorem [£3]. We 23 need only verify that F is a closed map, i.e., that if k. E K, k. 4 k, s. 6 F(k.) and s. 4 s then s 6 F(k). 1 1 1 1 1 Since each si is a minimizer Of fk , we have i fk (x) 2 fk.(si) for all x 6 K, 1 1 g(x)-f(ki)x‘2 g(si)--f(ki)si for all x E K. As i 4 m, we have by the continuity of f and g that g(x)-f(k)x 2 g(s)-f(k)s for all x 6 K, proving that s is a minimizer of fk’ In other words, 5 E F(k). Hence by Kakutani's theorem, F has a fixed—point, so that there exists a E K such that a 6 F(a). This says that g(x)-f(a)x 2 g(a)-f(a)a for all x 6 K, or equivalently g(x) 2 g(a)4—f(a)(x-a) for all x E K. Since 9 is infinite outside K, this characterizes f(a) as a subgradient of g at a. We introduce more assumptions regarding the supply model, so that the hypotheses of the PIES models are met. We then apply Theorem 2.2 to a sequence of approximations to arrive at Theorem 2.6. 24 2.3 Definition. We define the set Q to be eff dom v f) IRE Q [q20:qux for some x EX}, where X is given by X = {x.2 0: Ex g b], B and b as in (1.0.3). Q is the set of demand vectors of interest to us from the economic standpoint. We have shown as a consequence of Proposition 1.11 that eff dom v is closed, and so Q is closed. We assume that X is bounded, which is in accord with its economic interpretation. It follows that Q is also bounded, and so is compact. We further assume that Q is nonempty, and more specifically that the linear program (1.0.3) is feasible for some q 22 0. Under these assumptions, take K to be Q, g to be v, and f to be p in Theorem 2.2. The theorem then tells us that an equilibrium exists if p is continuous on Q. As noted earlier, this condition is not satisfied by the function p in the PIES model. In this case, p is undefined on part of bd Q, namely Q (1 bd IRE . The following lemmas will be used in the course of the proof of the next theorem. 2.4 Lemma. Let f be a convex function on 151, x a point at Which f is finite; then 2 6 af(x) iff f’(x;y) 2 zy for all y 6 Rd, 25 where f’(x;y) = lim t-l{f(x4-ty)-f(x)}. 1110 Proof: See Rockafellar [11]. 2.5 Lemma. Let v be as in (1.0.3) and Q as in Definition 2.3; then as q ranges over Q. av(q) assumes at most a finite number of subsets of I51. ‘nggg: As a consequence of Theorems 1.5 and 1.6, av(q) is the set Of all y 6 Iii for which there exists a vector 2 6 Hi: such that (y,z) is Optimal in the dual to (1.0.3); that is, av(q) is the projection into I51 of the set of Optimal solutions to the dual. The set of optimal solutions to the dual is a face of the polyhedrally convex set {(y,z) :y20, z 20, yA-ngc}. Polyhedrally convex sets have finitely many faces, proving the lemma. In the following theorem we use the subscripts i,j and k to denote the components of a vector and the sub- scripts m and n to denote the members Of a sequence. 2:6 Theorem. Let v be as in (1.0.3), 0 as in Definition 2.3, and p continuous with . d . d p.1nt 1R+ 4 1nt1R+. Assume that Q is compact and that p satisfies the 26 following conditions for i = l,...,d and q E Q: if pi(q) 4 +co then qi 4 0. (2.6.1) where pi and q1 are the ith components of p and q respectively; and if qi 4 0 then some pj(q) 4 +w. (2.6.2) Under these assumptions there exists an equilibrium demand q E Q, i.e., a vector q such that p(q) E av(q). Proof: Let e 6 3C1 be the vector (l,1,...,l). _ . . d . d For n — 1,2,... define pn..na_-4 1nt Ig_ by pn(q) = p> 0, it suffices to show that the set is empty. Suppose it is not. For i E I, 1 -1 e)i = qn,i + n 4 O as n 4 m, n 6 N. (qni-n In view of (2.6.2), 1 pn'j(qn) = pj(qn+n e) 4 +c=o (2.6.3) for some j (not necessarily i). For each k e {l,...,d}\I, qn'kz‘ 0 and so by (2.6.1) p (q ) = p (q +n'1e>/' +°° n,k n k n ° Thus for such k, some subsequence of (pk(qn)) is bounded. Hence we may extract a subsequence (qn)n€L from the subsequence (q ) such that for each n nEN k E {l,...,d}‘\I there exists a finite number xk such that ) 4 x as n 4 w, n E L. (2.6.4) pn,k(qn k By Lama 2.50 {6V(z) : z 6 Q} 28 is a finite collection of sets, and so we may pass to yet another subsequence, such that ov(qn) = ov(qm) for all m,n 6 M. We have postulated the existence of a vector y 6 Q such that y 2) 0; since qn,i 4 O for i 6 I, n 4 w, n 6 M, surely we can find a number m 6 M such that q < yi for all i 6 I and all n 6 M, n'2 m. (2.6.5) n,i Since v is convex, v'(qm:y-qm).g V(y)-V(qm) < ”- Since pn(qn) E 6V(qn) Bv(qm) for all n 6 M, n > m, by Lemma 2.4 pn(qn)(y-qm) g V’(qm:y-qm) < w for all n 6 M, n > m. Hence lim sup pn(qn)(y-qm) g V’(qm:y-qm) < w. (2.6.6) n-Doo n6M Now pn(qn) (y - qm) :21 pn'kmn) (yk - qm'k) + izejl pn'i(qn)(yi"q.m'i)r (2.6.7) 29 noting that a sum taken over an empty index set is zero. For k E I, Pn.k(qn)‘yk‘qm.k) " xk(Yk’qm,k) as n " °°' n 6 M' and so the first sum in (2.6.7) converges to a finite limit as n 4 m. For i 6 I, yi"qm,i > O by (2.6.5) and pn'imn) > 0. so the second sum in (2.6.7) consists exclusively of positive terms. By (2.6.3), there exists a j such that Pn j(qn) 4 +co as n 4 m, n 6 M; I in view Of (2.6.1), that j must belong to I, and so pmmnnyj-qm) * +°° as n * °°' Thus the second sum in (2.6.7) diverges to +m as n 4 m, n 6 M, contradicting (2.6.6). The theorem is proved. Conditions (2.6.1) and (2.6.2) are not satisfied by every function p of the form (1.0.1), so they represent nontrivial restrictions for the PIES model in particular. Since q is allowed to move only in a bounded set, (2.6.2) will hold if, for instance, the matrix M in (1.0.2) is nonpositive with negative diagonal, a condition satisfied at least by the example published in Nissen and Knapp [Si]. Condition (2.6.1) is an economic consequence if p 3O realistically represents the behavior of actual consumers, since violation Of (2.6.1) requires an infinite supply of money in the economy. The following example shows that (2.6.1), or some similar condition, must be satisfied. 2.7 Example. Let d = 2, r = 1, s = 2, A: (<1) ‘1’). B: (11).b= (1). C: (1.1) and p(q) - (qilqgl.q;1) for q >> 0. (2.7.1) Here Q=£q20=q1+nglh Observe that the inverse of p(-) is q(p) = (pilp2.pgl): if we let pl 4 +co and maintain p2 = p11, we see that ql 4 l, violating (2.6.1). In searching for an equili- brium, we need only consider q 2) 0, since p is undefined for q1 = O or q2 = O. For q )2 O, we Obtain by direct calculation: {(1.1)} if q1+q2< l. ql >0. qz > 0; OV(q) = (2.7.2) (12 > 0. Now from ql > O and (2.7.1), we infer that p1(q) 7‘ p2(q) for q 6 Q. q >> 0. 31 and so from (2.7.2) we see that P(g) E 5V(Q) for all q E Q. q >> 0. i.e., no equilibrium exists. We close the chapter by noting that the use of equality demand constraints in the PIES version Of (1.0.3), while presenting difficulties mentioned previously, foreshadows the following observation: at an equilibrium, the most economical production program must meet all demands exactly. We prove this below. 2.8 Proposition. Let v be given by (1.0.3), let . d . d p.1nt IR+ 4 1nt 1R+ be continuous, and suppose that there exists a vector q such that p(q) E av(q). Let i be an Optimal solution of (1.0.3) for that q; then Ax = q. nggf: Note that for p(q) to be defined, we must have q 2) 0. Also, the optimal solution i must exist, since av(q) #’¢ implies v(q) is finite which in turn implies that the linear program is feasible and bounded. If the ith demand is satisfied with slack, i.e., if (Ax)i > qi, by the complementary slackness principle of linear programming the corresponding component of any optimal solution to the dual of (1.0.3) must be zero. As was shown 32 in Chapter I, there exists a vector 2 such that (p(q),z) is Optimal in the dual to (1.0.3), and so Since the range of p lies by assumption in the interior of the positive orthant, we must have Ax=q. Note that this result is not surprising, as we have legislated it by requiring that both the argument and the value of p(-) be strictly positive. CHAPTER III UNIQUENESS OF EQUILIBRIA In this chapter, we address the question of whether an equilibrium for our models, if one exists, must be unique. We require a monotonicity condition on p, a higher-dimensional generalization Of the idea that, for one commodity, demand is a decreasing function of price. We make the following definition. 3.1 Definition. Let X C an , f :X 4 Ian . We say f is strictly monotonically decreasing iff (f(x)-f(y))(x-y) < O for all x,y 6 X, x #y. We now prove uniqueness Of the equilibrium When p is strictly decreasing. 3.2 Theorem. If x,y 6 Q, p(x) 6 av(x), p(y) 6 av(y) and p is strictly monotonically decreasing on Q, then X = y. Proof: Since p(x) 6 av(x), v(y) 2v(x)+p(x)(y—X). (3.2.1) 33 34 Similarly, since p(y) 6 Bv(y), V(x) 2V(y)+p(y)(x-y). (3.2.2) Adding (3.2.1) and (3.2.2), we have 0.2 (P(X)-p(y))(y-x). Since p is strictly monotonically decreasing, O< (p(x)-p(y))(y-x) if x7‘Y. and so x must equal y. Sweeney [15], among others, has observed that under the assumption that p' is globally negative definite, the equilibrium, if one exists, is unique. Note that here and in the sequel we apply the terms negative definite and negative semidefinite to asymmetric as well as symme- tric matrices. An arbitrary n> 0. Since G is Open and f’ is continuous at y, there exists e > 0 such that 36 hf’(y4-th)h > O for all t 6 [-e,c]. Again using continuity Of f’, ‘we have 6 j hf’(y+th)h dt > o; 0 but (6 hf’(y+th)h dt h(f(y+ch) -f(y)) o %((y+eh) —y)(f(y+ch) -f(Y)) < 0. a contradiction. This completes the proof. Monotonicity Of p is central to the proof of unique- ness of the equilibrium. In the PIES model, monotonicity is a consequence Of the negative definiteness Of p', as shown above. The authors of the PIES model made negative definiteness of p’ a standing assumption.[ 5]. we shall produce arguments to the effect that, for the log—linear form of p, this assumption is exceedingly restrictive when enforced globally, and will exhibit an example with multiple equilibria in which p’ is not globally negative definite. We begin with the case d = 2. 3.4 Lemma. For p given by (1.0.1) with d = 2, if p'(q) is negative definite for all q >2 0, either m12 = m = 0 (3.4.1) 21 or 37 4-1, m = m 4-1 and m = m 21 11 12 22 (3.4.2) m i-m 11 224-1 < 0. Proof: We remark first that a necessary condition fer a 2)<2 matrix H to be negative definite is that t % det(H4-Ht) = det(§J%E—) > 0. (3.4.3) We compute p’: m -l m m m -l 11 12 11 12 k1m11q1 qz klm12ql q2 p’(q) = . m —1 m m m -l 21 22 21 22 k2m21ql qz k2m22ql q2 Let q = (3,1), 5 > 0; then (3.4.3), with H p’, becomes m -m +1 1 l 2 2 ll 21 k1k2(m11m22"2 m12m21) ’ Z k1m12S - -l 1 2 2 m21 m11 m11+m21—1 after division by s . For this inequality to hold both as s 4 +w and as s 4 O, we require that either (3.4.1) or (1,5) shows that hold. A similar analysis with q either (3.4.1) or must hold. Assuming that (3.4.1) fails to hold, we can reduce (3.4.4) to 38 O < k k m m l 2 1 2 11 22 ‘ Z [kl(m22+l)+k2(mll+l)] _ - 1 _ 2 ‘ "k1k2(m11+’“22+1) 4 [k1(m22+1) k2(m11+l)] ' and so p’(q) is negative definite only if m + m + 1 < 0, ll 22 completing the proof. For higher dimensions, global negative definiteness is even more restrictive. 3.5 Proposition. For p as in (1.0.1), d > 2, if p’(q) is negative definite for all q >2 0 then for all i # j, either lj = mji = 0 (3.5.1) or m.. = m..4-1, m.. = m..4—l, and 13 33 31 11 (3.5.2) m..4-m..+-1 < O. 11 3] In addition, if (3.5.1) fails, then for all k such that i 71 k 7‘ :1. Proof: Assume that p'(q) is negative definite for all q 2) O. This in particular requires that the 2 x 2 submatrix .. d.. 11 13 (3.5.4) d.. d.. 31 J] 39 Of p’ = (dm). formed by the intersection of rows i and j with columns 1 and j in p’ (i # j), must be negative definite, and so (3.5.1) and (3.5.2) follow from Lemma 3.4, taking qk = l for i #’k # j. Now suppose (3.5.1) fails to hold for some i # j, and select h such that i #‘h #’j. Let qk = 1 for all k #’h, qh = s > O, and examine the submatrix (3.5.4), which reduces to I k.m..s 1h k.m..s 1 11 1 1] m. m. k.m..s 3h k.m..s 3h . 3 31 J 33 Condition (3.4.3) for this matrix becomes m. +m. m k.k.m..m..s 1h 3h - l [k.m..s lh-tk.m..s 3h]2 > 0, 131133 4 113 331 or equivalently m. —m. m. -m. 4k.k.m..m.. > k?m.?s 1h 3h-I-2k.k.m. m .-tk?m?.s 3h 1h, 1 j 11 33 1 13 1 j 13 31 j 31 Since one of mij'mji is nonzero, this cannot hold for all s 6 (O,+m) unless (3.5.3) holds. In economic terms, mij is the elasticity of price i with respect to demand j, a measure of consumer re- action to prices. Proposition 3.5 places exceptionally tight restrictions on what these elasticities could be. In practice, the demand model is constructed to fit actual consumer behavior, and it is quite unlikely that the conclusions of PrOposition 3.5 would be satisfied, assuming 40 the specified commodities bear some relation to each other. We are thus led to abandon the assumption that p'(q) is negative definite for all q 22 O. we might ask that p’(q) be negative definite at least for q 6 Q, but if the proofs of Lemma 3.4 and Proposition 3.5 are revised to allow s to go to O (as it certainly may in practice as q moves in Q) but not to go to +m, we arrive at more complicated but still highly restrictive necessary conditions. Hogan [5 ] describes negative definiteness Of p' as "a weak economic assumption and an Observed property of the PIES demand functions over the relevant regions," but does not explicitly describe the "relevant regions." Proposition 3.3 indicates that we will be hard pressed to relax the assumption Of negative definiteness significantly while retaining monotonicity. Our next example shows that without monotonicity, we may well have multiple equilibria. 3.6 Example. Let d = 2, r = 1, s = 2, A = (i i). B = (1 1). b = (2). c = (1,1) and l 2 -2 -1 p(q) = (32q1 q; .32ql q2 ) for q >> 0. We assert that (4,2), (2,4) and (3,3) are equilibrium values for q. Observe that p(4,2) (201): ML!“ (1.2). 41 and p(3.3) = (32/27,32/27). The dual to problem (1.0.3) is Zzl+z2 2 4-22 -w‘g 1 l 2 (3.6.1) zl'ZZ'W‘Z 0 -w g_l qlzl-tq222-2w (max) The triples (2,1,4), (1,2,4) and (32/27,32/27,23/9) are Optimal in (3.6.1) when q is (4,2), (2,4) and (3,3) respectively, which, in light of Theorems 1.5 and 1.6, proves the assertion. CHAPTER IV THE PIES ALGORITHM In the remainder of this paper we will assume the existence of at least one equilibrium for our models, and pursue methods for locating an equilibrium. The previous chapter indicated that multiple equilibria are a possi- bility When p is not strictly monotonic. From the standpoint Of utilization of such models as tools in policy making, we should be able to find all equilibria, but as yet we cannot. Hogan and others sought equilibria under the assumption that p’(q) is negative definite for each q. We make the comparable assumption that p is strictly decreasing, ensuring that the model has a unique equilibrium. With an eye toward simplifying future calculations, and with Proposition 2.8 in hand, we revise (1.0.3) using equality demand constraints, Obtaining Ax = q Bx g_b (4.0.1) x 2.0 cx (min). 42 43 'k We define the set Q , corresponding to Q in Definition 2.3, to be * d Q = A(X) n 1R+ 1 (4.0.2) where X is as in Definition 2.3. Moreover, we will assume, unless otherwise stated, that p is strictly monotonically decreasing, so that the equilibrium is unique. Lastly, we must face the fact that p(-) is not defined on all of 0*. We have two options. We could work on a compact subset of Q* bounded away from the coordinate hyperplanes, or we could assume p(-) is defined on all of 0*. The former alternative is plausible, since equilibria are strictly positive vectors, but we have no a priori knowledge Of where to truncate Q*, and in addition, we would be adding constraints to (4.0.1). We therefore choose the latter alternative. The PIES algorithm is motivated by the observation that if p has a potential f, i.e., if there exists a function f on 0 such that p = Vf, and if all the other assumptions above are met, then q is an equilibrium demand if and only if q minimizes v-f. 4.1 Lemma. If f and g are prOper convex functions and there exists a 6 rel int (eff dom f) D rel int (eff dom g), then fi-g is convex and 44 a(f+ g) (X) = of(X) + 69(X) for all x 6 eff dom f n eff dom 9. Proof: See Rockafellar [11]. We remark that for S,T C I51, the expression S-tT denotes the vector sum {s-tt :s 6 S, t 6 T]. 4.2 Proposition. If p is strictly monotonically * decreasing on Q and f : IR? 4 IR is a potential for p, i.e., vf(x) = p(x) for all x 6 int R3, then q is an equilibrium iff q minimizes v-f. Proof: Assume that the hypotheses hold, and let F = v-f. Since p is strictly monotonically decreasing, * * -f is convex, and so F is convex on Q . Since Q d *- is a convex set and F = +w on IH_\ Q , F is convex on JR:i . From Definition 1.4, q is an equilibrium iff p(q) e av(q). (4.2.1) Since f is finite on If?, F is proper, and so q minimizes F iff O 6 oF(q). (4.2.2) We need only show the equivalence of (4.2.1) and (4.2.2). We have assumed that there exists a vector y 6 Q such that y 22 0; without loss of generality y 6 Q* and so rel int (eff dom v) n rel int (eff dom -f) # ¢. Using Lemma 4.1, 45 aF(q) = av(q)+a(-f)(q) = av(q) -p(q) * for all q 6 Q . Thus 0 6 aF(q) iff P(Q) 6 av(q). proving the proposition. Necessary and sufficient conditions for the potential f Of a Cl function p on R3 to exist are that api/qu = apj/aqi for all i,j = l,...,d. Hogan [5 ] notes that these conditions are not met by the PIES indirect demand function p, and so p is not integrable independently of path, which implies that p has no potential. We will nonetheless be guided at times by Proposition 4.2. The computationally easiest case to handle is when each component function pi is a function only of the corresponding variable qi, i.e., p(q) = (91(q1).....qd(qd)). This occurs precisely when p’(q) is a diagonal matrix for all q (with negative diagonal entries, since p is strictly decreasing). Proposition 4.2 applies here, and in fact the potential f is given by d qi f(q) = k + Z) [a qi(t)dt i=1 i for an arbitrary constant k and arbitrary a 6 int IRE . Wagner [17] describes a method in which the problem Of minimizing F, as in Proposition 4.2, can be approximately solved by solving a linear program.which is an expansion of (4.0.1). We will elucidate this now. 46 Let I I be closed intervals in [O,+m) such 1'...'d that the rectangle I = Id x...xIa_ contains both a and the unique equilibrium point q. Partition each interval Ij into a finite number of subintervals with partition points w ...< w. 3'0 <...< w. j.-n < am such that wj O = aj. There is no need to use the same I number of points in each interval or even to use equal number of points on either side of w, but it simpli- 3:0' fies the notation to do so. Let f . a gl(wi,J) J — O'ooO'n gi,j=( > I lzl’ooo'd _ w. ' = _n'...'_l L gl( 1'3) 3 J and Ailj = wirj'l'l-wigj' J = -n'...'n-l' l = 1I°°Oldo The following result appears in Wagner's paper. 4.3 Lemma. The Optimal solution (x,y) to the linear program Bx glb n-l -l (AX)i = ai + J§OY1,j-jZ:J-n yl’j' l - l,...,d x.2 0 (4.3.1) 0 g'yirj S-Aipj’ l = 1'.."d’ J = -nloooon-l d n-l cx — Z) . . . . min / i=1 jELn 91,3yi,3 ( ) exists and satisfies the following properties for i = l,...,d: 47 1f yi,m 2 O for some m 2 0, then yi,j = O for all j < O and yi,j = Ai,j for j: O'ooo'm—l; (403.2) 1f yi,m 2 O for some m < 0, then Yi,j = O for all j 2_O and yi,j = Ai,j for j = m+1,...,-1. (4.3.3) In addition, the component ”1 corresponding tO demand constraint i in any Optimal solution to the dual program satisfies one of the following: 0 < Yi,j < Ai,jo j 2.0. vi = qi,j: (4.3.4) 0 < Yi,j < Ai,j' j < 0. Ii = -qi,j: (4.3.5) yi,j = Ai,j' Yi,j+1 = 0. i 2.0. (4.3.6) 910' 2'”i 2-9i,j+1’ Yi.j = Ai,j' Yi,j_1 = 0: y < 0. (4.3.7) '9i.j-1 2 "i 2 ”91.3” Remarks. The key to the proof is that each component function gi is monotonically decreasing and positive. The partition essentially estimates the integral Of each gi (i.e., the area under the graph Of gi) by a sum Of rectangles, taking into account the sign reversal when integrating from right to left. Letting nil _ “g. _ z. = a. + y. . - y. ., 1 1 j=O 1,3 j=-n 1,3 the sum n—l _ .2 91 . jyi .j 48 is just a Riemann-type sum for z. fa: gi(t)dt. (4.3.2) and (4.3.3) state that the approximation by rectangles is consistent with the integral, i.e., that rectangles do not occur on both sides of ai and that no gaps appear between rectangles. The proof consists of showing that when this is not the case, a redistribution of weight among the variables yi j can be made to reduce the value of ' n-l the Objective function while preserving the value of Z) y. .. . 1,3 =-n (4.3.4)-(4.3.7) are proved by characterizing an Optimal dual solution as a subgradient of the value Of (4.3.1) and then observing the effect on the Objective function Of a perturbation in the most extreme nonzero yi j for each i. I 4.4 Notation. To connect problem (4.3.1) to the problem Ax = z Bx g_b (4.4.1) x 2.0 d z. v(z)-f(z) = cx-f(a) - Z} Ial gi(t)dt (min), i=1 i we establish a correspondence between vectors (x,y) feasible in (4.3.1) and vectors (x,z) feasible in (4.4.1) with z 6 I, via the following: “’1 '21 Z. = a. + Z y. o - y. 0. (4.4.2) 1 1 j=0 1,3 j=-n 1,] Each y clearly produces a z feasible for (4.4.1); each 2 produces a y feasible for (4.3.1) if we also require 49 that y satisfy (4.3.2) and (4.3.3) as well as the constraints 0 g,yi j S'Ai j for all i,j. 4.5 Lemma. Let (x,y) be feasible in (4.3.1) and let z be given by (4.4.2); then (x,z) is feasible in (4.4.1) and the values of the Objective functions of (4.3.1) and (4.4.1), evaluated at those x,y and z, differ by no more than nt4m [max Ai,j][gi(wi,—n)'-gi(wi, )1. (4.5.1) 113 “ Proof: Feasibility of (x,z) in (4.4.1) is clear. Since each gi is strictly decreasing, w. .+y. . 1:] 1:] 91(Wi.j)yi.jwafL j gimdt 2 91(wi.j+1)yi.j for i = l,...,d and j 2,0, and so wi .+yi . (Din/3. '3 gi(t)dt-gi(wi i,j .j+1’Yi.j A similar estimate holds for j < 0. If we add these inequalities as j runs from -n to n-l and over- estimate the right side by replacing y. . with max A. , 1:3 k 10k the right side teleSCOpes, and after summing over i we arrive at (4.5.1). We remark in (4.5.1) that the value Of gj at the endpoints Of Ij are independent of the mesh size, and so once the rectangle I is established, the accuracy of the value Of (4.3.1) in estimating the value Of (4.4.1) depends on the mesh sizes max A. . (i = l,...,d) with 1,3 3 (4.5.1) providing an explicit estimate. 50 A more critical issue is the accuracy Of the Optimal § Of (4.3.1) in estimating the Optimal z Of (4.4.1). The following proposition shows that arbitrary accuracy between § and 2 can be Obtained by taking a sufficiently fine partition, but the crucial estimate relies on a number which we cannot calculate in practice. 4.6 Proposition. Let (x,y) be Optimal in (4.3.1), let 2 correspond to § as in (4.4.2), and let (§,E) be optimal in (4.4.1). Given 6 > 0, there exists a 6 2 0 such that lE-§l < e if max Ai j < 6. i,j ' Proof: We denote by (M) the problem Obtained from (4.4.1) by adding the constraint (2.212 e. where e is assumed to be small enough such that (M) has a feasible solution. Let the minimal values of (4.4.1) and (M) be C and D respectively. Let G = D-C; o > 0 since all vectors feasible in (M) are feasible in (4.4.1) and any Optimal solution (x,z) of (4.4.1) must have 2 = E and hence cannot be feasible in (M). According to Lemma 4.5, there is a 6 2 0 such that if max Ai . < 6, then for every (x,y) feasible in i,j '3 (4.3.1), the value produced by (x,y) in the Objective function of (4.3.1) and the value produced by the correspond- ing (x,z) in the Objective function of (4.4.1) differ by at most 0/3. It follows that the Optimal value E of 51 (4.3.1) differs from the Optimal value C of (4.4.1) by at most 0/3, and in particular C-E2-o/3. (4.6.1) Now suppose we choose the mesh size to be less than 5 and still find that [24%) 26; then (32.2) is feasible in (M). Problems (M) and (4.4.1) have the same objective function. Let F be the value of that function at (§,2). Since D is the optimal value of (M), F 2 D = C+o. (4.6.2) From our choice of 6 and the observation that (§,§) produces the value E in (4.3.1), E-F2-o/3. (4.6.3) Adding (4.6.1)-(4.6.3). C 2_C+-o/3, which contradicts o 2 0. Note that the choice of 6 depends, through 0, on D, which is unknown in practice. Proposition 4.6 guarantees however, that if we were tO repeatedly solve (4.3.1) with mesh sizes tending to O, 2 would converge to 3. Before moving to the more general case, we note that if the partitioned rectangle I fails to contain the Optimal solution of (4.4.1), the vector 2 corresponding to the y Optimal in (4.3.1) will lie on bd I. In this eventuality, we may partition a new rectangle about 2 and repeat the process. 52 For the more general case, in which p’ is not even symmetric, let alone diagonal, the PIES algorithm attempts to exploit problem (4.3.1) by approximating p(-) by a function with diagonal derivative matrix. We now present the algorithm. 4.7 PIES Algorithm. Start with a vector pO 6 R3. k = Or e E [001]- Step 1: Calculate q = p-1(pk). Step 2: Define an approximation g(';q) to p(°) by the following rule: 9i‘w’q) = pi(ql"'"qi-l'wi'qi+l'°"'qd) for i = l,...,d. Step 3: Solve approximately problem (4.4.1), taking gi(°) to be gi(-;q), by solving (4.3.1), Obtaining a solution (x,z), and set Wk = g(z;q)° Step 4: If Wk = pk (to within predetermined tolerances), stOp with (approximate) equilibrium demand q. Otherwise, set pk+l = 9Pk+ (1‘9” ' increment k by 1, and repeat from Step 1. Both Hogan [5] and Wagner [17] report that the algorithm typically converges within ten, and Often within six, iterations. Wagner states that termination occurs when the maximum deviation (presumably in p) is within 1%. 53 He further states, and HOgan concurs, that taking 9 =1/2 rather than 0 accelerates convergence. Despite these findings, the algorithm has to date evaded complete analysis, due in part to the fact that the function defined in Step 2 does not approximate p(°) in a fashion that we find tractable. The results we have on this algorithm are fragmentary. We first introduce another concept from convex analysis. * 4.8 Definition. The conjugate v of a convex function v : IRd4 [—co,+oo] is given by * d d v (x) = sup[xy-v(y) :y 6 It } for all x 6 ll . We note that since v(y) = +co for y 6 eff dom v, w v (x) = sup[xy-v(y) :y 6 eff dom v}. When v is the value of (1.0.3) or (4.0.1), we have the following result. * 4.9 Lemma. v is a proper convex function and *‘k v = (v ) . Proof: Since v is both lower semicontinuous and prOper, the lemma follows from a result in Rockafellar [11]. 4.10 Lemma. y 6 av(x) iff x 6 av*(y). Proof: See Rockafellar [11]. We note that ov*(p) is precisely the set Of demands q for which p is a possible supply price. 54 In all subsequent analysis we will assume that in Step 3 of the PIES algorithm, problem (4.4.1) is solved exactly. We denote its Optimal solution 2 by z = Z(q) to show the dependence on q. :k * . ' 4.11 Lemma. Z :Q 4 Q Is continuous. Proof: We note first that gi(-;°) is jointly continuous for each i. Let g = (gl....,gd). Since g’(';q) is negative definite, there is a unique 2 Optimal in (4.4.1), and so z = Z(q) iff g(z;q) 6 ov(z). Let q 6 0* and suppose (qn) C 0* such that qn 4 q. Let yn = Z(qn) 6 0*. Since 0* is compact, (yn) clusters at some y 6 0*. Passing to a subsequence, we may assume yn 4 y. By the closedness of the point-to- set map av(-), we have g(y;q) 6 av(y) and so y = Z(q). This holds for any cluster point of the original sequence * (yn), so by compactness of Q , Z(qn) 4 Z(q), i.e., Z is continuous. 4.12 Prgposition. Let g be an equilibrium demand, _ _ w - p = p(q). If v is differentiable at p, then the PIES algorithm converges to equilibrium in one step when started in a suitable neighborhood Of 5. Proof: We begin by noting that since ov(q) is closed for all q and only finitely many sets av(q) exist [Lemma 2.5], there exists a neighborhood M of 55 5 such that for all q 6 0*. 5 (Z ov(q) implies av(q) n M = (5. (4.12.1) Since g(-;-) is jointly continuous and g(g;g) = p(g) = 5 6 M, there exists a neighborhood N Of a such that y,z 6 N implies g(y;z) 6 M. Since Z is con- tinuous and Z(g) = 6 [because g(g;g) = 5 6 av(g)], we can find a neighborhood V of a such that V C N and Z(V fl 0*) C N. Suppose that we start the algorithm with pO 6 p(V), noting that p(V) is a neighborhood of B. Let q0 = p-1(po). y = Z(qo). qO 6 V, so qO and y belong to N and thus g(y;qo) 6 M. Since g(y;qo) 6 ov(y), av(y) D M.# ¢, and so by (4.12.1) E 6 av(y), which implies by Lemma 4.10 that y 6 av*(§). Since v* is differentiable at 5 and g 6 av*(§). av*(§) = {5]. Thus y = g. '1'. _. _ We remark without proof that av (p) = {q} when q is an extreme point of Q and p 6 int av(g). * 4.13 Lemma. For x,y 6 av (z), v(x)-v(y) = z(x-y). * Proof: Using Lemma 4.10, x and y in 6v (2) implies that z 6 av(x) and z 6 av(y), and so v(y) 2 v(x)4—z(y-—x) and v(x) 2_v(y)-+z(x-y). Combining these inequalities proves the lemma. 56 * 4.14 Lemma. 6v(x) = {y} for all x 6 int av (y). * Proof: If int av (y) = ¢, dbere is nothing to prove. * Suppose x 6 int av (y). Let u 6 Iii be arbitrary. * There exists t > 0 such that xi-tu 6 av (y), and so by Lemma 4.13 v(xi—tu)-—v(x) = tyu. (4.14.1) On the other hand, for any w 6 av(x) v(xi-tu)-v(x) 2_twu. (4.14.2) Combining (4.14.1) and (4.14.2), (y—w)u 2 O for all u 6 Rd, and so y = w. Since w 6 6v(x) was arbitrary, 6v(x) = {y}. 4.15 Proposition. If g is an equilibrium, p = p(g) _ * - and q 6 int av (p), then the PIES algorithm with e < 1 converges to 6 when started in a suitable neighborhood of p. _ *. Proof: Let N = Z 1 (int av (p)), a neighborhood of 5 since q 6 int 6v*(p), Z(q) = q and Z is continuous by Lemma 4.11. Since p and p.1 are both continuous on int Its, they are homeomorphisms there, and so we can find a ball M about 5 such that p-1(M) C N. Suppose that at some iteration k Of the PIES algorithm, pk 6 M; then q = p-1(pk) 6 N and so * _ _ y = Z(q) 6 int 5V (p), By Lemma 4.14. BV(Y) = (P)- 57 Since ”k = g(y:q) 6 av(y) = {5}. we must have Wk = 6. Thus )pk+l-§I = lepk+ (l- 9)7Tk"P) elpk-is Either pk = 5 (in which case the algorithm terminates) or lpk+l-—p[ < [pk-p‘. In the latter case, pk+l 6 M and the argument can be repeated, so that by induction — m — [pk-HH-p’ = 9 )pk-p) 4 0 We now present an example in which the indirect demand function, while always possessing a negative definite derivative matrix, contains a parameter we are free to set. For large values of the parameter, the PIES algorithm has failed to converge in over 100 itera- tions, and is believed by us not to converge at all. We shall give a heuristic argument for this and describe the Observed behavior. 4.16 Example. Let d = 2, r = l, s = 2, T 2 l, 2 l A = (1 2). B = (l l), b = (2). C = (1:1): With p defined by (1.0.2) with -T-1 -T 3(T4-l)log 2 M'= ( -T -T-1)' K = ((3T4-l)log 2)' 58 We note that the example is constructed to have the unique equilibrium q = (402): E) = (201): *... - and so that the hypothesis 6v (p) = [q] Of Proposition 4.12 holds. Our heuristic argument against convergence of the PIES algorithm on this example when T is large proceeds as follows. Suppose that the sequence (pk) generated by the algorithm converges to E, and so qk = p-1(pk)"&. Recall from the proof of PrOposition 4.12 that when ov*(§) = [a] and q is sufficiently near 5, Z(q) = a. Thus for k large, Z(qk) = q and so ”k = 9(Z(qk):qk) = g(aqu). Let _ -T-l O _ O -T . Md ‘ ) c) -T-1)' Mo ‘ (-T o)' then log wk = K-I-Md log q-I-MO log qk and log 5 = K+Md log g+MO log a, and so log Wk-log 5 = M0(log qk-log g). Now M is invertible, so from (1.0.2) log q = M-1 log p-M-1K 59 and hence _ _ -1 - log qk-log q — M (log pk-loq p). Therefore _ _ -1 _ - log nk-log p - MOM (log pk log p). . ' -1 T The eigenvalues of MOM are -T and 2T+-1 . write log pk-log p as a linear combination of the If we eigenvectors of M M-l, we arrive at log wk-log 5 O by stretching one component by a factor of -T and shrinking the other by a factor of '§é%fi, which approaches 1/2 as T grows large. If a were 0 (so that pk+l = wk), this would prevent log pk-log E from converging to 0 unless by some chance log pk-1og 5 were an eigenvector 1 of M M- The situation when ._;£__ 0 2T+1' e 2 O is less clear; we present this only to motivate the with eigenvalue choice of the particular example. A modification of the PIES algorithm was tried on the example, using a program written and executed on a Tektronix 4051 microcomputer, which holds numbers to fourteen decimal digit accuracy. The modification occurred in Step 3. Rather than using the linear program (4.3.1), we solved (4.4.1) directly, using the method of steepest descents. It was expected that, if anything, this would be more accurate than the use of the linear program approximation. 60 When run with e = 0.5 and T either 1 or 2, the sequence appeared to be converging to 5, even from fairly poor starts. When run with e = 0.5 and T = 3, however, convergence did not occur. Table 1 shows the distance (in the euclidean norm) between pk and 5 at various points during a number Of runs. In the first run, the program was started close to 5. After a few iterations, it began to eXhibit oscillatory behavior, appearing to be approaching two distinct limit points. To test this hypothesis, the program was restarted at what appeared to be one of the two limit points. As eXpected, the algorithm oscillated between (1.9999995, 1.00000025) P and (2.0000005, 0.99999975) P with the norm of the error constant to within machine tolerances. To test the effect of roundoff errors, the program was next started exactly at the solution. Ideally, the sequence generated should be identically 5, but in practice we would anticipate errors due to the compu- tation of logarithms and eXponentials and the finite accuracy of the machine. After ten iterations, the error was less than 3> 0 such that b 2 g(a;X)(b-a) -f 9(srx) - ds 2 m(X) lb-al (5.5.1) a for all a,b 6 0*. Proof: We note first that the integral on the left hand side of (5.5.1) is well-defined, since g’(o;x) diagonal implies that its integral is independent of the path. Fix a,x 6 0*. Define G : [0,1] 4 ]R by t G(t) = tg(a;x)(b-a)-I g(Tb+ (l-T)a;x)(b-a)d'r. Clearly O G(O) = O and G'(O) = g(a;x)(b-—a)-g(a;x)(b-—a) = 00 Moreover, G”(t) = —(b-a)g’(tb4—(l-t)a;x)(b-a) for 0 < t < 1. Using the second Mean Value Theorem, G(l) = -(b-a)g'(5b+ (l- 5)a:X)(b-a) for some g 6 (0,1). Since g'(-;x) is negative definite and continuous and Q* is compact, —g’(-:x) is uniformly positive definite on 0*, i.e. there exists m(x) > 0 such that -ug'(w;x)u 2 m(x)]u]2 for all w 6 Q* and all u 6 Rd. Thus b g(a;X)(b-a) -( g(S:X) - ds = 6(1) a = -(b—a)g’(gb+ (l- §)a;x)(b—a) ‘2 m(x)]b-—a|2 . 72 5.6 Proposition. Suppose that p 6 C1 such that g(-;-) is monotone in the first argument and satisfies the following Lipschitz condition: lg(x;y)-—g(x;z)| g_k ly-—z] for all x,y,z 6 Q*. Let ‘g be an equilibrium. If 1 < m(g), then the PIES-VAR sequence converges to g' from any start in q*. Proof: Since solves (4.4.1) at iteration qk+1 q _. v(qk+ 1) - a“ 1 g(s:qk) 0as g v(q) noting that the solution to (4.4.1) does not depend on what we take to be the lower limit of integration. Consequently _ q _ v(qk+ l) -v(q) —I_k+ l g(srq) 'ds q q (5.6.1) 3 3‘” [g(s;qk) -g(s;§)] °ds. q since 5 = 8(3) 6 ME) and 95:6”) = mg). v(qk+l)-v('g') 29(E:E)(qk+l-E) . (5.6.2) In view of (5.6.2) and Lemma 5.5, we see that ._ q ._ V(q )--v(q)-f__k+l g(s;q) °ds ki—l q _._. _. q ._ 2 g(QFQ)(qk+ 1-q)- __k+ l g(Siq) ' as q 2m(§)|qk+l-§lz . (5.6.3) 73 Now qk+1 __ J_ [g(Stqk)-g(s;q)] °ds q S -S—up lg(z;qk)—g(2;_q-)l iqk+l--q-l . (50604) Since lg(z;qk)-g(z;g)l g_k lqk-g], we have by combining (5.6.1), (5.6.3) and (5.6.4) that ma) lqu-Elzg) )qk-g) lqu-El and so lqk+l-' II p as in (1.0.2) with K 1.1 log 2-log 3 M -2 0-9 = 1.1 lo 2-1o 3 ' g g -1 —0.1 = (2 . 2). E: (1/3 . 1/3). .QI 74 This example is constructed with p’ negative definite throughout 0* (guaranteeing a unique equilibrium, namely a) and with '5 6 int av*(§) and av(g) = {5]. To verify this last claim, we note that for all q 6 Q*. the unique solution to (1.0.3) is x(q) = ((2q1-q2)/3, (-q14-2q2)/3) With V(q) = cx(q) = (qli-q2)/3. Thus {(l/3,l/3)} = {5). Moreover, for all p 6 R2 we have EVE) SUP (pq-V(q) Hi 6 R2} = sup (pq-V(q) :q 6 0*} v*(p) = SUP {pq- (ql+q2)/3 :q 6 0*} = sup ((p-Em =q 6 0*}. 0. Since ‘6 6 int Q*. and in particular v*(p) V*(p) -v*('§) = V*(p) .>_ (p-B)q for all p 6 JR2 and any q in a neighborhood of g in Q*, from which it follows that q 6 Bv*(§) for all q near 5' and hence 5'6 int av*(§). 5.8 Lemma. In example 5.7, g(§:q) 6 ov(g) implies that q =‘g. Proof: Since ov(g) = [E]. g(g;q) 6 Bv(g) implies that g(grq) = 5. NOW p(x) = (kxl—zxzo’g, kxl-lxz-O'l) for all x 22 0, where k = 3-121'1. Hence g(EIQ) = (k2-2q20'9,k2-O'lq1-1). since 3': (2,2). Therefore g(g;q) = 5' iff qzo'9 = 20'9 and q1-1 = 2-1. i.e. iff q=‘q'. 5.9 Proposition. If qO #“g, the PIES-VAR sequence (qk) cannot converge to q. 75 _P_r_o_c_>_f_: By Lemma 5.8, if qk+1 = (i then g(aqu) 6 OV(E) and so qk = 6. Since qO 316, it follows that qk 7! g for all k. Now suppose qk 4 6'6 int av*(p). By Lemma 4.14, av(q) = (P) for all q 6 int ov*(p). Since gk 4 a, there exists m > 0 such that qk 6 int av*(§) for all k 2 m. Using the decomposition M = Md+ MO' have g(qk+ l:qk) = p and so K+Md log qk+l+M0 log qk = for k2m we log 5. Also K+ Md log g-I-MO log 3 = log 5, and so for k2m _ _1 _ log qk+ 1- log q = -Md MO[log qk- 109 q] . It follows that for k 2 m - __, -1 2 — log qk+ 2 -log q — (Md MO) [log qk- log q] . Now -4.5 O (Md-l M0)2 = O -4.5 and so for k2m I log qk+2-logg I = 4.5 I log qk-log El. From this we deduce that for k = 1,2,... lloqq -1093) =4s" (1.. -10951 m+2k ' qm ' Since qk 4 q by assumption, I log qm+ 2k- log q l 4 0 as k 4 co, and so we must have q = a, a contradiction. m CHAPTER VI A SUBGRADIENT PROJECTION ALGORITHM The lack of theoretical justification for convergence of the PIES algorithm, and the inherent difficulties in analyzing it, have led a number of researchers, including Eaves [3] and Irwin [6], to prOpose other algorithms for locating equilibria in models of the PIES form. The failure of the PIES algorithm on example 4.16 underscores the need for a more generally convergent method. In this chapter we state an algorithm for minimizing a convex function of a particular type over a polytoPe. The algorithm and a proof of convergence were prOposed by Sreedharan. When the indirect demand function p has a concave potential f, the function v-f is of the specified type, and we will prove that our algorithm produces a sequence converging to the minimizer of v-f over 0*. The extremal point located is, by virtue of prOposition 4.2, an equilibrium. In addition, the function f is used as a tool to prove convergence but is never specifically evaluated by the algorithm, leaving Open the possibility that the algorithm will prove efficacious on some problems in Which p does not have a potential f. The problem of minimizing a convex function over a linearly constrained set has been the subject of much 76 77 study. We single out the approach taken by Rosen [12] where the objective function is smooth. Rosen attempts to exploit the known convergence of the method of steepest descent in the unconstrained case. As is typical of so-called "feasible direction" methods, Rosen computes at each iteration a direction of descent Which points into the constrained region from the current point. He searches in that direction until he reaches either a relative minimum along the ray of search or the boundary of the constrained region. The process then repeats. Rosen's contribution is the choice of direction. When possible, he uses the negative gradient as the direction; When this direction points out of the set, he projects it onto a face of the set. Rosen's method is susceptible to a phenomenon known variously as "jamming" or "zigzagging", in Which the sequence generated clusters at, or even converges to, nonOptimal points. The trouble lies in the possibility that the sequence is alternating among two or more faces of the constrained region in such a way that the distance along the direction of search from the current point to the boundary is going to zero. Various modifications have been prOposed to avoid this. In particular, Polak [10] has adOpted a technique which prohibits the sequence generated from approaching arbitrarily close to a face when conditions for a constrained minimum are not being met at the limit. 78 Rosen's method, and much of the other work in the area, requires that the objective function be differentiable. Even when p has a potential f, our objective function v-f is not differentiable everywhere because v is not. Attention has recently been focused on algorithms for Optimizing nondifferentiable convex functions. The algorithms of Wolfe [l9] and Lemarechal [ 7], which generalize classical methods for unconstrained Opti- mization by replacing the gradient with a carefully chosen subgradient, do not treat the constrained case. The algorithm of Bertsekas and Mitter [1.], which does handle constrained problems, requires computation of the "c-subdifferential" of the objective function, Which is prohibitive in the problem we consider here. The algorithnt proposed here is prompted by those of Sreedharan [13,14] , Rosen [12] and Polak [10]. It resembles the Bertsekas - Mitter algorithm but requires the computation of only a manageable portion of the c-subdifferential. In this chapter we pose the algorithm and prove its convergence; in the next chapter, we discuss the actual implementation and report on some trial applications. 6.1 Problem. Let X C IVE be a nonempty convex polyt0pe given by X = {x 6 Rd] aix g bi' i = l,...,m]. NOte that.polytOpes are, by definition, bounded. Let v.:]Rd 41R (j=l,...,r) be given by 3 d v.x= .x+c., .6]R,c.61R andletthe J() 93 3 93 J . 79 polyhedrally convex function v : Rd 4 R be given by v(x) = max{vj(x)l j = l,...,r}, x 6 Eta. (6.1.1) Let f : Rci 4 R be strictly convex and of class C1 on a neighborhood of X. Finally, denote by 6 the indicator function (cf. Rockafellar [11]) of X, i.e. 6(x) = 0 if x 6 X and 6(x) = +co if x K'X. We consider the problem ll H § 3 aix g b1' i . (6.1.2) f(x)-tv(x) (min) which is equivalent to the problem of locating an . . . . d unconstrained minimizer of F = f4-v4-6 over 11 . We note before proceeding that the problem of locating the equilibrium of the PIES model fits the form of (6.1.2) when the indirect demand function p has a strictly concave potential, for if we take X = 0* and take f to be the negative of the potential of p, then by theorem 1.8 the value v Of the linear program (4.0.1) is pothedrally convex on Q*, and by proposition 4.2 q is an equilibrium iff q solves (6.1.2). We introduce some needed notation. 6.2 Notation. For x 6 X and 6‘2 0, we define the following sets of indices: 80 I€(x) = (l g,i g_m) aiX-2 bi-e): (6.2.1) J€(x) = {l ng g_r| vj(x) 2_v(x)-c}. (6.2.2) Note that 10(x) = {1 g,i g_ml aix = bi} (6.2.3) and JO(X) = {l g_j g,r| vj(x) = v(x)}. (6.2.4) We also define two convex subsets of 151, namely Cc(x) = cone{ail i e I€(x)} (6.2.5) v and K€(x) = conv{gjl j 6 J€(x)}, (6.2.6) where for any set S we use cone S and conv S to denote respectively the convex cone, with apex at the origin, generated by S and the convex hull of S. For any nonempty closed convex set S C 151 there is a unique point x 6 S nearest to the origin, which we denote by N[S]. The point a = N[S] is characterized by the following: a = N[S] iff a(x-a) 2.0 for all x 6 S. (6.2.7) We now present a subgradient projection algorithm for problem (6.1.2). 81 6.3 Algorithm. Begin with arbitrary x0 6 X, co 2 0 and with k = 0. Step 1: Compute y0 = N[vf(xk)i—Ko(xk)-tco(xk)]. If y0 = 0, stop: x solves (6.1.2). If yO #’0, k set 8 80. Step 2: Compute y8 = N[vf(xk)4-K€(xk)-tC€(xk)]. 2 _ _ Step 3. If lye) 2 8, set 6k — 8, 5k — y€ and go to step 5. Step 4: Replace c with 8/2 and go to step 2. Step 5: Compute Gk = max{o 2 0| xkndsk 6 X]. (It will be shown that 6% 2 0). Find Gk 6 [0.3%] such that there exists 2k 6 vf(xk-dksk)4-K0(xk-dksk) satisfying zksk = 0; if no such Gk exists, set ak=ak° Step 6: Set xk+1 = xk-dksk, increase k by l, and go to step 1. The implementation of steps 1 and 2, which can be treated as quadratic programs, and of step 5, Which requires a special line search procedure, are discussed in the next chapter. We next state a sequence of lemmas leading to a proof that algorithm 6.3 solves problem (6.1.2), or 82 equivalently locates an unconstrained minimizer of F = f4-v4-6. ‘We must first define the "c-subdifferential." 6.4 Definition. Let G : Rn 4 [49,00] be a convex function. The e-subdifferential of G at the point x, denoted a€G(x), is defined by a€G(x) = (u e n“ [G(y) 2 G(x)+u(y-x) -e for all y e Rn}. The usual subdifferential 5G(x) of G at x is just BOG(x). We now state a sequence of lemmas, using the earlier notation. \ 6.5 Lemma. For all e 2_0 and all x 6 351, K€(x) C o€v(x). Proof: If u 6 K€(x), then by (6.2.6) there exist Xj-Z 0, j 6 J€(x) such that Z:lj = l and u = :3 1.9.. j6J€(X) 3 3 For j 6 J€(X), we have vj(y) = vj(X)4-gj(y-X) 2_v(x)-—c-+gj(y-x). (6.5.1) Therefore for every j 6 J€(x), v(y) = max vi(y) 2 v(x)-c-+g.(y-x), i=1,...,r 3 and so 83 v(y) = 2'3 X.V(y) 2 V(X)-e+ '23 Rely-X) j6J€(x) 3 j6J€(x) 3 3 (6.5.2) =V(X)-e+u(y-X). Since (6.5.2) holds for all y 6 I51, the lemma is proved. 6.6 Lemma. Given x 6 151, there exists a neighborhood V of x such that JO(y) C JO(x) for all y 6 V. gpppfi: The functions wj = v-vj, j = l,...,r are continuous with wj(x) 2 0 iff j z JO(X). Thus there exists a neighborhood V Of x such that wj is positive throughout V for each j g JO(X). If j g J0(x) and y 6 V, wj(y) 2 0, and so j z J0(y), proving the lemma. 6.7 Definition. Let S C 191 be a nonempty set. The support function P of S is defined by Cp(X) = SUP(XYI y 6 S}. x 6 Rn. 6.8 Lemma. Two closed convex subsets of 151 are identical iff their support functions are identical. Proof: See Rockafellar [11]. 6.9 Lemma. 6v(x) = KO(x) for every x 6 X. Proof: Let x 6 X. Since both av(x) and K0(x) are closed and convex, it suffices, in view of lemma 6.8, to ShOW’that they have the same support function, namely 84 I I v (x;-) [v defined as in lemma 2.4]. It is well-known that since v is everywhere finite—valued, v'(x;y) = sup {yul u e av(x)} for all y 6 Rd, (6.9.1) i.e. v’(x;-) is the support function of av(x). Given y 6 151, by lemma 6.6 there exists an e 2 0 such that JO(x+-dy) C J0(x) for all a 6 [0,6]. Since for 0 g_d g_e v(x4—oy) = max v.(x4-ay) j6JO(x+ay) and V(X) = max V-(X)r j6J0(x) 3 we see that v(xi-dy)-v(x) = max [v.(x4—oy)-—v.(x)] ' J 36J0(x) = _max dgjy. 36J0(x) This shows that v’(x;y) = rmnc g.y = max [uyl u 6 Ko(x)}, (6.9.2) - J 36J0(X) and so v'(x;-) is also the support function of KO(X). We note that (6.9.2) proves the following statement: 6.10 Corollary. v’(x;s) = max {sul u 6 K0(x)}. 6.11 Lemma. For each x 6 X and e 2 0 there exists a y > 0 such that JO(x) C J€(y) whenever lx-y| < y. 85 Proof. Choose y 2 0 such that lgjly < %- for j = l,...,r and [v(x)-v(y)] <-% if Ix-yl < y. Now if j6 J0(x) and [x-yl < y, then V(y) - vj(y) = My) -V(X) +vj(x) -Vj(y) (X‘Y)l < 60 <-§-+lqJ and so j 6 J€(y). 6.12 Lemma. oF(x) = vf(x)4-Ko(x)-tC (x) for all x 6 X. 0 ugpppg: The indicator function 6 is clearly prOper and convex, while f and v are everywhere finite valued. It is well-known that for x 6 X, 66(x) = CO(x). Moreover, any a 6 rel int X belongs to rel int (eff dom f) n rel int (eff dom v) n rel int (eff dom 6). The result now follows from lemma 4.1. The next lemma shows that the stOpping criterion in step 1 of the algorithm is well chosen. 6.13 Lemma. If y0 = O in step 1 of algorithm 6.3, then xk is the minimizer of F. Proof: y0 = 0 implies that 0 6 aF(xk), a necessary and sufficient condition for xk to minimize F. The strict convexity of f ensures that the minimizer of F is unique. 6.14 Lemma. Step 4 of algorithm 6.3 is not executed infinitehy often in any one iteration. 86 Proof: If step 4 is executed infinitely often, then 3 4 O and ye 4 0. Now 61 2 62.2 0 implies that K€ (xk) C K€ (x and ) 2 l k C€2(Xk) C C€1(Xk) and hence that [Y lily l. 61 ‘32 so that y€ 4 O as c 4 0 implies that y6 = 0 for every 6.2 0; but then y0 = O, and we cannot have reached step 4, a contradiction. We now show the practicability of step 5 of the algorithm. 6.15 Lemma. If sk # 0, then -sk is a feasible direction of strict descent at the point xk. Proof: From the definition of s in step 3 of k the algorithm, sk = N[vf(xk) + K€k(xk) +C€k(xk)] . Let 1 6 10(xk) C I€ (Xk); then ai 6 C (x k 6k k) and so 5 -+ai 6 vf(xk)+K€ k (xk)+C€ (xk). k k using the fact that Ce (Xk) is a convex cone. By k (6.2.7) we have sk(sk4-ai-sk)‘2 0. Thus aisk 2 0 for every 1 6 Io(xk). Since aixk < bi for i z 10(xk), 87 there eXists d 2 0 such that ai(xk-dsk) S-bi for all i = l,...,r. Hence -sk is a feasible direction at Xk. To show that -s k is a direction of strict descent, we show that F'(xk;-sk) = :13 [F(xk-dsk)-F(xk)]/o < 0. (6.15.1) From the first part of the proof, there exists ‘5 2 0 such 6 X for 0 g_d g/d. For a in this range, that x -as k k F(Xk-Gsk) = f(xk-csk)4-v(xk-dsk) and so by corollary 6.10 F’(Xk:-sk) = f’(X—k;—sk) +V’(xk:-sk) -vf(xk)sk4-max{-skyl y 6 KO(xk)} = -min((vf(xk)4-y)skl y 6 KO(Xk)]. (6.15.2) When y 6 Ko(xk) C Kc (Xk) we have k vf(xk) + y 6 Vf(Xk) + K€k(xk) +C€k(xk) and so by (6.2.7) sk(vf(xk)4-y-sk).2 0 and consequently (vf(xk)+y)sk 2|sk|2 > 0. Combining this with (6.15.2), we have F’(Xk7_sk)-§ -lsk|2 < 0, (6.15.3) completing the proof. From the first half of lemma 6.15 we have the following corollary. 88 6.16 Corollary. The number a; defined in step 5 of algorithm 6.3 is positive. The next lemma shows that in the relevant case the vector 2k in step 5 of the algorithm exists. 6.17 Lemma. Let sk # 0 and define m on (0,6k] by m(d) = F(Xk-Gsk). If 5% is not a minimizer of T on [0, k], then 2k satisfying step 5 of algorithm 6.3 exists. ggppfi: By lemma 6.15, m'(0) = F’(xk;-sk) < 0, so that there is some a 6 (0,5k] such that p(q) < m(O). Since we have hypothesized that 6k does not minimize cp, there exists Gk 6 (0,5k) minimizing m over [0,ak]. Set y = xk-dksk. There exists 8 2 0 such that F(y) g F(y+lsk) for [11 g c. It follows that [F(y+).sk)-F(y)]/). 20 (6.17.1) and [F(y-lsk)—F(y)]A 20. (6.17.2) 0 < l g.c. Since F is convex, the directional derivatives F’(y;sk) and F’(y;-sk) both exist, and from (6.17.1) and (6.17.2) we conclude that F'(y;sk) 2 0 and F'(y;—sk) 2_0. Using corollary 6.10, F'(y;:tsk) f’(y;i:sk)+v'(y;:ts k) ivf(y)sk+max{ :tuskl u 6 KO(y)]. 89 Since K0(y) is compact, there exist u,w 6 Ko(y) such that ._ ’ . Vf(y)sk-I-usk — F (y,sk) 2 O and __’ .- vf(y)sk+wsk — F (y, sk) g_0, and so for an apprOpriately chosen convex combination h of u and w we have h 6 K0(y) and vf(y)sk'+hsk = 0. Taking zk = vf(y)-th 6 vf(y)+-Ko(y) satisfies the requirement in step 5 Of the algorithm. The number 0k determined in step 5 of algorithm 6.3 has the following prOperty. 6.18 Lemma. Let sk # 0 and p be as in the previous lemma. Then Gk is the unique minimizer of p on [0,Ei]. Moreover, 0k is positive. gpppg: Since F’(xk;-sk) < 0 by lemma 6.15, the conclusion that Gk 2 0 follows immediately once we show that Gk minimizes m over [0.6%]. Uniqueness of this minimizer follows from the strict convexity of F. If zk satisfying step 5 of the algoirthm cannot be found, then by lemma 6.17 5% minimizes p over 90 [0,Ek], and in step 5 we set Gk = 3%. Suppose then that Gk 6 (0,3k] is located such that an apprOpriate vector zk exists. Set y = xk-dksk. Since zk 6 vf(y)+-Ko(y) C BF(y), for any a 6 [0.Ek] we have by the subgradient inequality that cp(c) = F(xk-csk) 2 F(y)+ ( 2 — 6km y‘s/2 — N[Vf(XkI)+K£(Xkl)+C£(xka)]. 2 2 From these we see that Iy2 c [123 28k), showing that k YZE: I 4 0. We now repeat the proof of lemma 6.21, k 93 replacing s with yze: 7' concluding that x 4 x. I k k 6.23 Lemma. The sequence (sk) is bounded. Proof: Note that Vf(xk) + K0(xk) C vf(xk) + K€k(xk) + C€k(xk) . so that (SR) 3 lvf(xk)l + 1N1K0(xk)11. (6.23.1) K0(Xk) is one of a finite number of possible polytOpes, so that there is an upper bound on [N[K0(xk)]l inde- pendent of k. As f is of class C1 on the compact set X, the right hand side of (6.23.1) is bounded, proving the lemma. 6.24 Lemma. If the sequence (sk) is bounded away from 0, then (0k) converges to 0. Proof: Suppose that (s is bounded away k) from O and that Gk f 0. Since dklskl is bounded above by the diameter of X and (sk) is bounded away from 0, (0k) is bounded. Given this and the compactness of X, we can pass to corresponding subsequences (Sk')' (0k.) and (xk.) such that Sk' 4 S # 0, 0k. 4 d 2 0 and Xk' 4 x 6 X. By lemma 6.19, the sequence (F(Xk')) is monotone decreasing, so that all of its subsequences have the same limit, namely F(x). In particular, F(xk'+1) 4 F(x); 94 but Xk’+l = Xk’ ”ak’sk' 4 x-ds , so that Fm-os)==FM). (624A) Since F is convex and F(xka-ak.sk.) g F(xk:-Ask:) for all A 6 [0.6%.], we have F(Xk' was.” s F(xx' - skew/2) s “X.” and so in the limit F(x-ds) gF(x-o(s/2) gF(x). (6.24.2) Since a 2 O and s #’0, (6.24.1) and (6.24.2) taken together contradict the strict convexity of F. 6.25 Lemma. Let the sequence ) defined in algorithm (ck 6.3 be such that there exists 6.2 0 satisfying ck.2 c for every k. For any index i, the inequality binaixk g_c (6.25.1) implies the inequality bimaixk ng._--aixk_*_l . (6.25.2) Proof: If (6.25.1) holds, then i 6 I (xk), and €k so a. 6 C (xk). As was noted in the proof of lemma 6.15, it follows that aisk 2_0. Since xk+1 = xk-dksk, (6.25.2) must hold. 6.26 Lemma. Assume that the following hold. (1) The sequence (ck) in algorithm 6.3 is such that there exists 6 2 0 with ek‘2 c for all k. 95 (ii) The sequence (0k) converges to 0. (iii) Some subsequence (Xk') of (xk) converges to the point x. Then there exists a subsequence of (xk.), again denoted (xk.), such that Io(xk.) = IO(X) for every index k'. ggppfi: Assume that (i), (ii) and (iii) hold. Since the index sets 10(xk.) are subsets of the finite set {l,...,m}, we can pass to a subsequence of (Xk’)' again denoted (xk.), such that for some subset I of {l,...,m] we have 10(xk.) = I for all k'. We must show that IO(x) = I. If i 6 I, then I = bi for all k’, so that in the limit aix = b.. a.x i k 1 Therefore I C Io(x). Now suppose that i 6 Io(x)\\I. We derive a contradiction. Since xk+1 = xk-dksk, with (s ) shown bounded in lemma 6.23 and Gk 4 0, k we see that lxk+l-xkl 4 0 as k 4 m. Hence there exists k such that O ai(xk+l-xk) < %- for all k 2_k0. (6.26.1) Choose p 2kO such that 10(xp) = I and 6* = bi"aixp < 8. (6.26.2) Such an index p exists because i 6 IO(x) implies that bi-aixk. 4 bi-aix = 0. Also 6* 2 0, Since i f I. Let q be the first index such that q 2 p and bi-aixq g .*/2 . (6.26.3) 96 Now by (6.26.1). (6.26.2) and (6.26.3) bi"ain-l = bi-aixq4-ai(xq-xq_1) O. (6.28.10) 100 On the other hand, letting k’ 4 w in (6.28.6) yields 25 = 0, (6.28.11) contradicting (6.28.10). Thus our assumption that (xk) fails to converge to the solution of (6.1.2) cannot be valid. The proof that the algorithm generates a sequence converging to the optimal solution is now complete. CHAPTER VII IMPLEMENTATION OF THE SUBGRADIENT PROJECTION ALGORITHM In this chapter we propose methods for implementing algorithm 6.4 and describe computational experiences. The algorithm was coded in FORTRAN on a CDC 6500 computer and tested on several problems of the form described in section 6.1. We present here the results of those tests. The program used a value of 10'.5 for the parameter co in algorithm 6.4. For computational efficiency, we divided 8 by 10 rather than by 2 in step 4 of the algorithm. Step 4 can be rewritten so that e is replaced by e/a for any fixed a 2 l. The algorithm terminated when the euclidean length of the projection yO in step 1 was less than a specified figure, usually 10-10. The two major problems in implementation were the nearest-point projection subroutine, required in steps 1 and 2 of the algorithm, and the line search procedure used in step 5. We first describe the line search. In step 5, it is required to determine a 6 [0,5] such that 25 = O for some 2 6 vf(x-as)4-KO(x-ds). Using the notation established in section 6.3, the orthogonality condition becomes 23 Bigis + vf(x-ds)s = O , (7.0.1) i6JO(x-ds) 8120 forall i, 23 B.=l. i6JO(x-ds) lOl 102 Rewriting (7.0.1) as Z) Bi[g.s + vf(x-ds)s] = O , (7.0.2) i6JO(x-ds) l we note that a solution (Bi) exists iff for i6JO(x-ds) some i,j 6 J x-—ds) either gis-tvf(x-ds)s = O or O( gis+-vf(x-—ds)s and gjs4-vf(x-ds)s have Opposite signs. We can summarize the characterization of a as follows. 7.1 Lemma. 0 6 [0,5] is the desired solution in step 5 of algorithm 6.4 if and only if either (a) there exist i,j 6 JO(X-ds) such that [915+vf(X-qs)S][qu+vf(x-qs)S] g0 (7.1.1) or (b) d = a and no triple (0,1,j) with O g_d £23 and i,j 6 J (x-as) satisfies (7.1.1). O Our line search proceeds as follows. We begin with d = 0 and locate the first value 3 2 a at which JO(x-as)\JO(x-ds) # q. We check whether there exist i,j 6 JO(x-3s) satisfying (7.1.1). If not, we check whether there exists i 6 J (x-as) n J0(x-ds) such 0 ~ A that gis4-vf(x-ds)s and gis-tvf(x-os)s have Opposite signs. If so, we locate the value of 0 between 6 and A a for which gis4-vf(x-ds) = O and terminate the search. This value of a can be found by Newton's method When f is 103 A ._ sufficiently smooth. If no such 1 exists and 0 < 0, ~ A we replace a with d and repeat the process. 7.2 Line searchAprocedure. Begin with x,s,a given. Step 0: Set a = O and consider all indices in Jo(x) as untested. Step 1: Choose an untested index i 6 JO(x-ds). A 1.. Step 2: Compute d = max {0 6 [0,0] :vi(x-os) = A a. v(x-ds)]. If a = 0, go to step 1. A Step 3: For each j 6 JO(x-ds) do the following: 2;: If A [gis4-vf(x-ds)s][gjs-tvf(x-as)s] g_0 , A terminate the search with a = d; .gp: If j 6 JO(x-ds) and .. A [gjs+-vf(x-ds)s][ngA-vf(x-ds)s] < O , go to step 5. A — ~ , A Step 4: If a < d, replace 0 With a and go to step 1, treating all indices as untested. Otherwise, terminate the search with 0 =‘E. ~ A Step 5: Locate the value a 6 [0,0] for which gjs4-vf(x-ds)s = O and terminate the search. 104 7.3 Projection procedure. The other major problem in implementing the algorithm is the calculation of the nearest point to the origin in a convex set. The convex set in question is vf(x)4-K€(x)-tC€(x) with c 2_0. For convenience, let uj = gji-vf(x), j = l,...,r. Since vf(x)4-K€(x) = conv {uj :j 6 J€(x)}, the projection problem can be posed as: ( h k W 061R+.fi361R+ k v 4.15:]. 5 < v=l v (7.3.1) 1 h k 2 3 2: 0a:.L + Z Bu.l (min) 0:1 H p v=l V 3v J where I€(x) = {il""'ih} and J€(x) = [j1,...,jk}. The objective function is a quadratic form in 0 and B, and so the projection problem is simply a quadratic programming problem. Letting g = (0,6) and H = (a. ...a. u. ...u. ) 1l 1h 31 3k ' we can write the objective function as 1/2 5 HtH 6, where (using superscript t to denote transposes) GtG GtU ! I UtG UtU / I G = (a. ...a. ) and U = (u. ...u. ). 1 1h 11 HtH is positive semidefinite but not positive definite, and problem (7.3.1) can in general have multiple optimal solutions. Despite this, there exists a unique nearest 105 point N[vf(x)4-K€(x)-tC€(x)], since vf(x)4-K€(x)4-C€(x) is a convex set and the euclidean norm is a strictly convex norm. Our approach is to locate a vector 6 = (0,5) satisfying the Kuhn-Tucker necessary and sufficient conditions for Optimality in the quadratic programming problem (7.3.1). These conditions are: ( §.y6m:'+k.w613 e§ = l g ( t (7.3.2) H H§-y4-we = 0 L 6y = 0 2 h k . where e= (0,...,0)x(l,...,1) 6R xR and w is a sign-unrestricted scalar. We initially employed the first phase of the Dantzig Two-Phase linear programming procedure [12], with a restricted basis-entry rule to maintain the complementarity condition gy = O. This method can encounter nonOptimal tableaus in which further pivoting is blocked by the complementarity restriction, even when HtH is positive definite. A Simple modification, however, solves this. Since all equality constraints, with one exception, are homogeneous, we can construct an initial Phase I - feasible solution by setting any one of the BV equal to unity and completing the basis 106 with artificial variables (which may require multiplication of some of the homogeneous equations by -l). From this point onward, the modified Phase I procedure reduces to an algorithm of Wolfe [18]. Our elimination of all linear terms in the objective function, by means of the substitution uj = vf(x)4-gj, causes our problem to satisfy conditions under which Wolfe's proof of convergence of the algorithm applies with only minor alterations, even through HtH is only semidefinite. Wolfe assumes that the system of equality constraints is nondegenerate, i.e. that the constraint equations are linearly independent and that no basic feasible solutions with a basic variable equal to zero occur. The problem before us clearly satisfies the assumption of independent constraints, while Wolfe's proof can be modified to obviate the assumption that no basic variable vanishes. 7.4 Computational results: the PIES counterexample. Algorithm 6.4 was first tested on example 4.16, although convergence in this case has not been proved due to the lack of a potential for -p. When tested with the parameter T = 3 and a convergence criterion of less than 10-10 error in the euclidean norm of q, the algorithm converged in one to three iterations from a variety of starts. The general pattern was one step from the starting point to an edge of Q* containing the equilibrium vector (4,2), and then another step to 107 the solution. Where a third step was required, it appeared to be a very short step, perhaps correcting rounding errors. The only difficulties occurred when the initial point was too close to the origin, where we believe the size of the components of -p(Q) is SO much greater than that of the generators ai and gj as to cause severe roundoff errors. The algorithm was also tested with parameter values T = 7 and T = 10, with similar convergence results. 7.5 Computational results: the PIES-VAR counterexample. We next tested algorithm 6.4 on example 5.7. AS with the previous example, the "gradient" -p does not actually have a potential, and so convergence of the algorithm is not guaranteed. Example 5.7 was attempted with three different starts. When started at q = (1,1), the algorithm reached the equilibrium (2,2), to within 10”8 euclidean norm, in one iteration. From a start at (4,2), convergence was oscillatory: the error was approximately 10 after 2>(lO-7 after 40 iterations and 2.2)(10- 60 iterations. Started from (1,2), however, the algorithm appeared to fall into a four-iteration cycle. unlike the previous attempt, in which the norm of the error decreased monotonically, the norm of the error in the four step cycle fluctuated between .511 and .961. 108 7.6 Computational results: Wolfe's example. Wolfe [19] considers an example in two variables with f identically zero: v(x) = max {V1,v2,v3] (min) where vl(x) = -x1, v2(x) = xl-I—x2 and v3(x) = xl-2x2. The level sets of v are triangles nested about the origin. The example was solved under the added constraints -10 S.le x2 g 10. The global minimizer (0,0), interior to the constraint region, was reached in at most two iterations from a variety of starts. Similar results were obtained using constraints which placed (0,0) on the boundary of the feasible region. The method also reached the constrained Optimum in at most two iterations using constraints which put (0,0) exterior to the feasible region. We note that Wolfe's example (and Powell's example, which follows) can be posed as linear programs, but with no gain in Speed of convergence. 7.7 Computational results: Powell's example. Wolfe [19] reports the following example due to Powell, on which the conjugate gradient method converges only linearly: v(x) max {vj(x) :j = O,...,4} (min) where vj(x) cos(2wj/5)4—xzsrn(2wj/5). Linear convergence x1 of the conjugate gradient method is Observed when started at any point of the form (p cos(Wj/5). P sin (Wj/5)). The 109 contours of v are regular pentagons centered at the minimizer (0,0). Algorithm 6.4 converged to the solution in two iterations from any feasible starting point using constraints which placed the origin in the interior of the constraint region. Similar results occurred when constraints were used which placed the origin on the boundary of the feasible region. USing the constraints x1 g_O, x2 g_0, xl-i-x2 g -l, for which the origin is infeasible, the algorithm typically took approximately five to ten iterations to reach the constrained minimum. 7.8 Computational results: a larger example. The algorithm was tested on an example having nine variables. The feasible region was a hypercube with the Optimal solution at one vertex. The smooth part f of the objective was a strictly convex quadratic, and there were three affine functions Vi' The algorithm converged to 13 within 10- of the solution (which was a unit vector), in at most ten iterations, from a variety of starts. LIST OF REFERENCES [l] [2] [3] [4] [5] [6] [7] [8] [9] [10] [ll] [12] LIST OF REFERENCES D.P. Bertsekas and S.K. Mitter, "A Descent Numerical Method for Optimization Problems with Nondifferentiable Cost Functionals," SIAM J. CONTROL, 11, 4 (1973). G. Dantzig, Linear Programmipg and Extensions, Princeton University Press, Princeton, NJ (1963). B. Curtis Eaves, "A Locally Quadratically Convergent Algorithm for Computing Stationary Points," TR SOL 78-13, Department of Operations Research, Stanford University (1978). W.W. Hogan, "Energy Policy Models for Project Independence," Comput. and Ops. Res., 2 (1975). W.W. Hogan, "Project Independence Evaluation System: Structure and Algorithms," Proc. p£_§ymposium.ip Appl. Math. 9; the A.M.S., 21 (1976). C.L. Irwin, "Analysis of a PIES-type Algorithm," presented at the Symposium on Energy Modeling and Net Energy Analysis, Colorado Springs (1978). C. Lemarechal, "An Extension of Davidon Methods to Nondifferentiable Problems," Math. Prog. Study 3 (1975). H. Nikaido, Convex Structures and Economic Theory, Academic Press, New York, NY (1968). D. Nissen and D. Knapp, "A Regional Model of Interfuel Substitution," Energy: Mathematics and Models, Proc. of a SIMS Conference on Energy. SIAM (1975). E. Polak, Computational Methods ip_0ptimization, Academic Press, New York, NY (1971). R.T. Rockafellar, Convex Analysis, Princeton University Press, Princeton, NJ (1970). J.B. Rosen, "The Gradient Projection Method for Nonlinear Programming, Part I," J. SIAM, 8, 1 (1960). 110 [l3] [14] [15] [l6] [17] [18] [19] 111 V.P. Sreedharan, "Least Squares Algorithms for Finding Solutions of Overdetermined Linear Equations which Minimize Error in an Abstract NOrm," Numer. Math.. 17 (1971). V.P. Sreedharan, "Least Squares Algorithms for Finding Solutions of Overdetermined Systems of Linear Equations which Minimize Error in a Smooth Strictly Convex Norm," .q. Approx. Thry.. 8, l (1973). J. Sweeney, "On the Uniqueness of PIES Solutions," Department of Engineering Economic Systems, Stanford University (1976). R.S. Varga, Matrix Iterative Analysis, Prentice-Hall, Inc., Englewood Cliffs, NJ (1962). M. Wagner, "Project Independence Evaluation System Integrating Model," Energy: Mathematics and Models, Proc. of a SIMS Conference on Energy, SIAM (1975). P. Wolfe, "The Simplex Method for Quadratic Programming," Econometrica, 27, 3 (1959). P. Wolfe, "A Method of Conjugate Subgradients for Minimizing NOndifferentiable Functions," Math. Proq. Study 3 (1975). "I(l((l'lllfllllttlfliis