BEFUWTSOK THEGRY NYE-'1‘ REPUSMENS ”EC GHEMQXL EEQCTEOH EQUAEGNS ' This is to certify that the thesis entitled BIFURCATION THEORY WITH APPLICATIONS TO CHEMICAL REACTION EQUATIONS presented by Ms . Nancy Theresa Waller has been accepted towards fulfillment of the requirements for P114). degree in Mathematics 'éL'wrxL UL Major professor . 0-7639 ? “3m av F 5- “W 0mm IND: LIBRARY BIND? ES :rmaun were 3“ ||||l|llllll|||||llllllllllllllllllllllllllllllllllllllllll :2”th 3 1293 10530 9474 ABSTRACT BIFURCATION THEORY WITH APPLICATIONS TO CHEMICAL REACTION EQUATIONS BY Nancy Theresa waller This thesis concerns families of nonlinear differential equations in a Banach space which depend on one or more para— meters. At certain critical values of the parameters, non- trivial equilibrium states may bifurcate from the trivial solution. ‘we consider two cases. In the first case, the generalized null space of the linear part of the system is one-dimensional at the bifur- cation point, and the system depends on a single parameter. We determine the number and magnitude of the bifurcating solutions and their stability properties. The second case involves dependence on two parameters. We consider the situation where there are two "bifurcation" curves in the parameter plane which intersect transversally. The linear part of the system which corresponds to these curves has a one-dimensional generalized null space, except at the intersection where it is two-dimensional. We develop analytical methods which can be applied to study the number and magnitude of the bifurcating solutions as a function of Nancy Theresa waller the parameters near the bifurcation point. We then apply these techniques to a system of partial differential equa- tions which arises in the study of chemical reactions. BIFURCATION THEORY WITH APPLICATIONS TO CHEMICAL REACTION EQUATIONS BY Nancy Theresa waller A DISSERTATION Submitted to Midhigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mathematics 1976 In memory of my grandmother, Amelia Zegaren ii ACKNOWLEDGEMENTS I am indebted to my thesis advisor, Professor Shui—Nee Chow, for his stimulating mathematical discussions. I am especially grateful to my academic advisors, Professors Shui-Nee Chow and Lee M. Sonneborn for their encouragement and the confidence they had in me. I extend my appreciation to the faculty of the Mathe- matics Department of Michigan State University, to the secretaries and my friends; and to my husband for his love and support. Finally, I would like to thank Ms. Jill Hagan for her time and her careful preparation of the final text. iii CHAPTER I: §1. 92. §3. §4. TABLE OF CONTENTS PRELIMINARIES . . . . Bifurcation . . . . . . . . . . . Differential Equations and Stability Analytic Functions and the Implicit Function Theorem . . . . . . . . . Remarks . . . . . . . . . . . CHAPTER II: BIFURCATION AND STABILITY-ONE §1. §2. §3. §4. DIMENSIONAL NULL SPACE . . . . Preliminaries . . . . . . . . . . . Reduction to the Finite-dimensional PrObleln O O O O O I O O O O O O O O Bifurcation [One Dimensional Null Space] . . . . . . . . . . . . . . Stability . . . . . . . . . . . . CHAPTER III: A TWO-DIMENSIONAL NULL SPACE g1. §2. g3. g4. §5. §6. Introduction . . . . . . . . . . . The Bifurcation Equations . . . . . An Appropriate Change of Scale Solutions of TYpe I-Fold curves Solutions of Type II. . . . . . Solutions near (XO’YO’YO’ iv 0) = (0,0,0,0). Page 12 15 23 30 3O 34 37 45 54 71 Page CHAPTER IV: APPLICATION TO A SYSTEM OF CHEMICAL REACTION EQUATIONS . . . . . . 77 §l. Introduction . . . . . . . . . . . . . . . 77 92. Eigenvalues, Eigenvectors, and Projections . . . . . . . . . . . . . . . 80 §3. Calculation of the Bifurcation Equation . . . . . . . . . . . . . . . . . 84 §4. Analysis of the Bifurcation Equation . . . 92 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . 104 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure 10. 11. 12. 13. 14. 15. LIST OF FIGURES Transversal crossing of zero eigencurves . . . . . . . . . . . . Dependence on (T,n) of the number of nontrivial solutions of (3.1) . . . . . Correspondence between (T,n)-plane and (n,Y)eplane . . . . . . . . . . . . . . Dependence on (n,y) of the number of nontrivial solutions of (3.2) . Behavior near a fold curve in the (§,y)-plane . . . . . . . . . . . . . . Behavior near a fold curve in the (T , 7") -plane a o o o o o o o o o o 0 Behavior near a cusp . . . . . . . . . Bifurcation diagram in the (T,Y)-plane . . . . . . . . . . . . . . Bifurcation diagram in the (T, n)-plane o o o o o o o o o o o o o 0 Behavior of nontrivial solutions of system (6.2) . . . . . . . . . . . . . The curve Red; = O . . . . . . A typical curve of neutral stability . The sectors S1 and S 2 O O I I O Bifurcation diagrams for the system (3.10) in the (T,n)-plane . . . . . . . Intuitive representation of bifurcation When he < O and c/a < O. . . . . . . vi Page 33 38 39 4O 53 53 68 69 69 72 82 82 92 101 102 CHAPTER I PRELIMINARIES §l. Bifurcation As a simple example of bifurcation, consider what is observed experimentally when a compressive axial thrust is applied to a thin elastic rod [12]. As the thrust T is gradually increased from zero, the rod first becomes thicker and shorter, but its center line remains straight. The classical linear theory of elasticity predicts this straight state to be the unique equilibrium state of the rod for all values of T. However, when T readhes a certain critical value To, the rod is observed to buckle into a bent state which becomes more pronounced as T in- creases. The classical theory is inadequate to describe the phenomenon of buckling. When nonlinear effects are no longer neglected, one may construct a model which predicts both the straight state and the bent state as possible equi- librium states for T greater than To. Intuitively, the straight state ”loses its stability" to the bent state at this critical value. Let X, Y, and A be Banach spaces, and let F: AxX-pY be continuous. We will say that x O is an equilibrium state corresponding to do in A if x0 is a solution of F(ao,x) = 0. we will say that (ao,x0) is a bifur- cation point for T if and only if (i) F(a0,x0) = 0 (ii) For every neighborhood V of (a ,xo), there exist a in A and.:xl,x2 in. X. with x1 # x2, such that (a,x1) and (a,x2) are in V, and F(a,xl) = F(a,x2) = 0. That is, we can always find a arbitrarily close to do such that there is more that one equilibrium state near xo that corresponds to a. §2. Differential Equations and Stability First we shall mention some standard results, from the theory of ordinary differential equations [£3]. Let A: £514 If} he a continuous linear map, and let N: If!» 151 be continuous, with N(O) = O, DN(O) = 0 [where DN(O) de- notes the Fréchet derivative of N at x = 0]. Consider the ordinary differential equation (2.1) %%=Ax+N(x) and the associated linear differential equation :13: (2.2) at = Ax. Definition 2.1. The zero solution of (2.1) (or of (2.2)) is said to be uniformly asymptotically stable if and only if there exists 6 > 0 such that lxbl < e implies |x(t)| 4 0 as t 4 m where x(t) is the solu- tion of (2.1) (or of (2.2)) with x(0) = x0. A necessary and sufficient condition that the system (2.2) be uniformly asymptotically stable is that all the eigenvalues of A have negative real parts. If this is the case, there exist positive constants, K and a, such that -a(t-t0) lx(t)\ < ‘xol Ke where x(t) is the solution of (2.2) with initial condition x(to) = x0. If N(x) is O(|x|2) as x approaches zero, the asymptotic stability of (2.2) implies that the zero solution of (2.1) is asymptotically stable. ‘We will now consider differential equations in a Banach space X. Let A be a closed linear operator whose domain is a dense subspace D(A) contained in x. It is useful to consider D(A) as a Banach space Y with the so-called graph norm: |z|Y = lzlx + |Azlx. The injection j: Y 4»X is continuous with dense range, and A: Y 41x is then a continuous linear map. Let N: Y 4.x. be continuous, with N(O) = O and DN(O) = 0. Consider the differential equa- tion (2.3) %% = Az+N(z) and the associated linear differential equation dz _ (2.4) 3-1:- — AZ. Here dz dt z(t+h)-z(t) lim hHO h where the limit is taken in the norm l-lx. Motivated by the results for ordinary differential equations, we make the following definition. Definition 2.2. we shall say that the system (2.4) is stable if there exists a 5 < 0 such that whenever A is in the spectrum of A, Rex < 5. We shall say that the system (2.4) is unstable if part of the spectrum lies to the right of the imaginary axis. In many cases, given that (2.4) is stable in this sense, it is possible to prove stability results analogous to those for ordinary differential equations [10], [l3], [14]. That is, if the linear system (2.4) is stable in the sense of definition (2.2), then the zero solution of the nonlinear system (2.3) is stable. These results depend on the particular properties of the operator A, and we will not go into them here. Instead we make the following hy- pothesis. Principle of Linearized Stability. Let 2 = 20 be an equilibrium solution of (2.3). If the linearization of (2.3) about 2 given by O) dz _ (2.5) at — Az+DN(zo)z is stable in the sense of definition (2.2), then the solu- tion 2 = z of (2.3) is stable. 0 In what follows, we shall say that 20 is stable, if (2.5) is stable in the sense of definition (2.2), and that z is unstable, if (2.5) is unstable. Statements about 0 stability of 2 will actually be statements about the 0 location of the spectrum of the operator A-PDH(zO). §3. Analytic Functions and the Implicit Function Theorem For the sake of completeness, we mention the following results, which may be found in Dieudonne's Foundation 2: Modern Analysis [ 5 ]. Definition 3.1. Let D be an open subset of KP, where K = I! or C. We say that a mapping f of D into a Banach space E over K is analytic if, for every point a e D, there is an open polycylinder P = [z E Kpllzi—ail<(ri, l g i g p}, such that in P, f(z) is equal to the sum of an absolutely summable power series in the p variables (Zk-ak), l < k < p. The fbllowing are true: (3.2) The power series in (3.1) is unique. (3.3) Let A.c Cp be an open connected set, f and 9 two analytic functions in A with values in a com- plex Banach space E. If there is a nonempty open subset U of A such that f(x) = g(x) in U, then f(x) = g(x) for every x in A. Let U be an open subset of A, b a point of U, and suppose that f(x) = g(x) in the set U{1(b:+lé5, then f(x) = g(x) for all x in A. (3.4) Let E be a complex Banadh space, A an open sub- set of fig), f an analytic mapping of A into E. Then there is an open set B c GP such that B nIIRP= A and an analytic mapping of B into E into which extends f. (3.5) A continuously (Fréchet) differentiable mapping f of an open subset of cp into a complex Banach space is analytic. [Henceforth, differentiable will mean Frechet differentiable]. Implicit Function Theorem. Let E, F, G be three Banach spaces, f a continuously differentiable mapping of an open subset A of ExF into G. Let (xo,yo) be a point of A such that f(xo,yo) = O and the partial de- rivative D2f(xo,yo) be a linear homeomorphism of F onto G. Then there is an open neighborhood U6 of XO in B such that, for every open connected neighborhood V of x0, contained in Ub, there is a unique continuous mapping V into F such that u(x0) = yo, (x,u(x)) e A and f(x,u(x)) = O for any x e V. Furthermore, u is contin- uously differentiable in V, and its derivative is given by (3.6) u(x) = -[D2f(x,u(x))]_1[le(x,u(x))]. If f is p times continuously differentiable in a neigh- borhood of (xo,yO differentiable in a neighborhood of x ), then u is p times continuously O . The fol lowing also hold (3.7) If E, F, G are finite dimensional and f is an- alytic in A, then u is analytic in a neighborhood of x0. (Here Act):p or AcRp). (3.8) If E = mp, then u: E 4.F is continuously differ- entiable, hence analytic by (3.5). §4. Remafiks In this thesis we will be interested in two special cases of the following problem which we briefly outline here. Given a family of differentiable equations Oslo rrN = A(a)z-+N(a,z) (as in 92), which depend on a parameter a in Cn, describe the set of equilibrium solutions near a bifurcation point (aO,O) of the operator F(a,z) = A(d)z-+N(a,z). In particular, we shall be interested in the number of real equilibrium solutions which correspond to a in is). ‘We shall intro- duce the hypotheses we need and make our notions more pre- cise in the chapters that follow. In Chapter II, we use the Liapunov-Schmidt method [7 ] to reduce the problem to a finite system of “bifurcation equations" on a finite dimensional space. We shall then specialize to a case where a is in C and generalized null space of A(ao) is one dimensional, and determine the set of bifurcating solutions along with their stability properties. This situation arises in fluid dynamics and.has been studied by Kirchggssner and Sorger [13] in the context of the Taylor problem, and by Kirchgassner and Kielhdfer [14] in a general survey of bifurcation in fluid dynamics. Sattinger [17] has used Leray-Schauder degree to study the stability of bifurcating solutions, and has obtained results whidh overlap those of Chapter II. The technique which we employ is different, and shows how the sign of the critical eigenvalue is related to the leading terms in the bifurca- tion equations. Chow, Hale, and Mallet-Paret [(4] have studied a two parameter bifurcation problem which concerns the buckling of a rectangular plate. In one of the situations they - studied, the generalized null space is two-dimensional. In Chapter III, we shall develop methods for analyzing the bi- furcation set in a different general setting Where the gen- eralized null space is two-dimensional and the system depends on two complex-valued parameters. ‘We apply these results to a system of chemical reaction equations in Chapter IV. CHAPTER II BIFURCATION AND STABILITY-ONE DIMENSIONAL NULL SPACE §l.- Preliminaries This chapter is divided into two parts. The first two sections are largely introductory: we consider the problem of determining nontrivial equilibrium states which bifurcate from the trivial solution of 23% = A(a)z +N(a,z), and give basic hypotheses under which this problem can be reduced to a finite-dimensional problem. In §3, we consider the case where the generalized null space of A(O) is one- dimensional. We shall assume that for real a, A(a) has a simple real eigenvalue x*(a) which crosses the imaginary axis as a moves through zero. The number of bifurcating solutions and their dependence upon a will be discussed. In g4, we shall study the stability properties of these solutions and show how stability is related to the leading terms of the bifurcation equation. We now give our basic hypotheses and discuss their con- sequences. 10 (H1) Let A be an open set in Tn, and let X be a complex Banach space with dense subspace D(A). Let A: A xD(A) 4 X be a closed linear operator with domain D(A) fer each a in A, and let A be analytic in A in the sense that A(a)z has a Taylor expansion at each a in A which converges in a disc la-—a I < r indepen- O 0 dent of 2. Assume the origin is in A. It follows from (H1), that if Y' is the Banach space consisting of D(A) endowed with the norm |z|Y = |z|x + [A(O)z[x, then for some neighborhood U of the origin in Tn, we may regard A: U)(Y-+X as a continuous map for each a in U. (H2) Let N: U)(Y-ox be a continuous map such that N(a,O) = O and D2N(a,0) = O for all a in U; [i.e., N has zero linear part at (a,O) ]. We assume that N is continuously Fréchet differentiable in a neighborhood V of (0,0). According to Nachbin [15], this implies that for every V0 in V, there is a p > O and a power series itfla 1 m ET'Qm‘V"Vo) that converges to N uniformly for [v-v0[ < p. D is a m symmetric m-linear form on [CnVXY]m. In fact, Dm is the 11 th m Fréchet derivative of N at v denoted by DmF(vO). 0) In most applications, this series is finite. In most physical problems, we deal with real spaces and real parameters. Thus we assume (H3) A and N are extensions of "real" operators in the sense that A(a)z = A(a)§ and N(a,z) = N(5,z). Now consider the differential equation (1.1) 3% = A(a)z+N(0L,z) and the steady state equation (1.2) O = A(a)z+N(a,z). The trivial solution is always an equilibrium state for a near a = 0. If A(O) is a linear homeomorphism of Y onto X, then the implicit function theorem guarantees that the only solution of (1.2) in a neighborhood of (0,0) is the trivial solution. Thus we may expect nontrivial equi- librium solutions to bifurcate from the trivial solution at (0,0) only if A(O) is not a linear homeomorphism; i.e., A = O is in the spectrum of A(O). We now assume (H4) A = O is an isolated eigenvalue of A(O) with finite dimensional generalized null space. For our purposes, we shall assume that the null space is equal to the gener- alized null space. 12 §2. Reduction to the Finite-dimensional Problem we shall show that the problem of determining equilib- rium solutions of (2.1) corresponding to a near a = O, can be reduced to a finite system of equations on the gen- eralized null space of A(O). In our case the generalized null space is the null space and is finite-dimensional. Let Px denote this null space. By standard results from theory of closed operators we have Proposition 2.1. There is a continuous projection P: XJ+P which commutes with A(O) in the sense that X A(O)Pz = PA(O)z for z in D(A) = Y. By means of this projection, X may be decomposed into two complementary subspaces, Px and Qx = range (I-P). Each element 2 in X can be written uniquely as a sum 2 = b-+w where b is in PX and w is in Qx. By setting P = erjY = P and QY = Qxij, we have a corresponding Y X decomposition for Y'c X. These spaces are invariant under A0 in the sense that A(O): PY-on Furthermore, the spectrum of the restriction of A(O) and A(O): QY-on. to QY does not contain A = O, and we have Proposition 2.2. The restriction A(O): QY-on is a linear homeomorphism. 13 The Liapunov-Schmidt method.['7] consists in using the projections P and I-P given by proposition (2.1) to decompose the equation (2.1) O = A(a)z+N(a,z) into an equivalent system of two equations as follows. Let z in Y’ be rewritten as 2 = w-+b where w is in QY and b is in PY; b may be considered as a point in Cm where m is the dimension of Py. Equation (1.2) is then equivalent to the system 0 (I - P) {A(O) + [A(a) — A(O)]](w +b) + (I - P)N(a,w +b) (2.2) 0 II P[A(O) + [A(a) -A(O)]}(w+b) +PN(a,w +b) Since P and (I-P) commute with A(O), we have r(2.3) (a) o Fl(a,b,w) = Ao(w) + (I -P)[A(a) -A(0)] (w+b) < + (I-P)N(a,w+b) [(2.3)(b) O F2(a,b,w) = P[A(a)-A(0)](w~+b)-+PN(a,w-bb) where Fl: Ux‘lflluxQY 4 QX and F2: Ux 0. Example (3.1). Let A(u) = L-+pB for u in C, where L and B are densely defined closed linear operators on a Banach space X, with D(L) c D(B). Suppose that L has a continuous compact inverse and that L-lB has a con- tinuous compact extension to all of X. The spectrum of A then consists of isolated eigenvalues with finite multipli- cities. Suppose A = O is a simple eigenvalue of A at “O #'O, with eigenfunction go. Then we have analytic ex- pansions for the eigenvalue Mu) = “(u-HO) +A2(H_HO)2+... and the eigenfunction (9(a) = (904-291((4-(40) Ham-(1&2 +--- which are valid, in a neighborhood of p = “0 [11]. we claim that (H6) is satisfied. . l -1 Since 0 = (L-—pOB)¢b, we have 0 = [ES I-L B]¢b, and go is an eigenfunction for L-lB corresponding to the simple eigenvalue SL-. As in proposition 2.1, there is a projection P: X » span[qo} that commutes with L-lB. Then A(u)w(u) = (L-HB)m(u), and x(u)PL-lcp(u) = PL’1 (L - (whom). 17 Differentiating with respect to p gives PL"l (L - “oBN’l + PL-chpO -1 -1 -1 _ -1 or AlPL m0 — L 890. . —l -1 Since 0 =(I-L B)cpO and m0 #'0, L Bmo #'0. Thus A17’0- Many of the linear operators encountered in mathematical physics fall into this category. The aim of this section and its successor is to prove the following Theorem 3.1. Let A and N satisfy (H1)-(H6). In addition, suppose that the remaining spectrum of A satisfies [ReA[ > 5, for some 5 > 0. Then one of the following occurs (i) There is an infinite number of real nontrivial solutions of (2.2) near the zero solution for a = O ["ver- tical bifurcation"] (ii) For each real a near 0, there is a real non- zero solution z(a) of (2.2) whidh may be expanded in a fractional power series about a = O. 2(a) is the only nontrivial solution such that zl(a) tends to zero with a. (iii) There are two real nonzero solutions zl(a) and 22(0) for (2.2) for a > O (a < O) and none for a < O (a > 0). These solutions may also be expanded in fractional power series. 18 Theorem 3.2. Let L(a) denote the linearization of (2.2) about zl(a) [or 22(a)] given by L(a) = A(a)-+D2N(a,zl(a)). Under the hypotheses of Theorem 3.1, the spectrum of L(a) remains near that of A(O). For a near zero there is a simple real eigenvalue A+(a); A+(a) and A*(a) have oppo— site signs in a neighborhood of a = 0. Corollary 3.3. If the spectrum of A(a) (with the exclusion of A*) satisfies ReA < 6 < O for a near zero and A*(a) < O fer real 0 < O, and there is no vertical bifurcation, then real nontrivial solutions are unstable for a < O and stable for a > O. Sattinger [17] has proven a similar result using Leray Schauder degree when the linear operator A(a) is given as in Example 3.1. Gavalas [£5] has also applied degree theory to the stability of bifurcating solutions. In some cases the stability for particular systems which arise in fluid mechan- ics has been determined by perturbation methods [13], [18], [19]. The proof we give here is different and directly re- lates the sign of A+(a) to the derivative of the bifurca— tion equation. 19 We now return to the analysis of the system ['(3.l)(a) 0 Fl(a,b,w) = A(O)w + (I -P)[A(oz.) -A(0) 1 (w +b) 4 + (I-PO)N(OL,w+b) [(3.1) (b) o F2(a,b,w) = P[A(a) -A(O)] (w+b) +PN(a,w+b) where Fl: de: xQY 4 QX’ F2: U x03 xQx 4 C. We have already shown that for (a,b) near (0,0), there is a unique analytic w(a,b) such that F1(a,b,w(a,b)) = 0. Let w(a,b) = bloc-r 2 L bolb-+bllab-+b20a -+ ---+bkzakb . Since w = O is a solution for (a,b) = (a,O), and w(a,b) is unique in a neighborhood of (0,0), we must have w(a,0) = 0. Thus bk0 = 0 for all k. We now determine blo = g—% (0,0). Since F(a,b,w(a,b)) = O and D2N(O,O) = O, implicit differentiation shows that Em = - - - as = A(O) ab (0,0) 0. Since b01 IS in Qy, ab (0,0) 0, and we have Proposition 3.1. There exists an analytic function w(a,b) which solves (3.1)(a) in a neighborhood of (0,0). This function has the form 2 w(a,b) = b ab+b0 b +O(a2b+b2a+b3). ll 2 Upon substitution of w(a,b) into (3.1)(b), the bi- furcation equation becomes 20 (3.2) O = F(a,b) = aPAlb-+aPA1w(a,b) + P[A(a)-aAl-Ao](w(a,b)-+b)-+PN(G,b'+W(G,b)): where PN(a,b-+w(a,b)) is analytic and has a power series expansion in a neighborhood of (a,b) = (0,0). Recall that N(a,0) a O and D2N(a,0) E O for all a in a neighborhood of zero. Thus D§N(a,0) = O and D§D2N(a,0) = O for all a near zero, and we have PN(a,b-+w(a,b)) = b2N1(a,b) where N1(a,b) is analytic in a neighborhood of (0,0). It is useful to calculate the term PAlb. Let m(a) = eb-+aml-+a2m2-+-o- be the eigenfunction corresponding to x*(a) = 00.1 +a2A2 +--- Then P[A*(a)I-A(a)]m(a) = O. Differentiating with respect to a at a = 0 gives P[AlI-Al]¢b-+P[-Ao]ml = O or Almo = PAlqO. Thus PAlb = Alb. The bifurcation equation then becomes _ 2 o — Alab+aPOA1[b11ab+b02ab +... ] + P [azA -+a3A -+---](w(a b)-+b) 0 2 3 ’ 2 Thus a and b are related by a power series in two vari- ables with real coefficients: 21 b-+c a2-+0(ab-+b2)]. (3.3) O = b[Aloz+clo 02 The trivial solution b = O is a solution for all a near 1 2 . Clo to be EPDZN (0,0) \ Cpo,¢po> where wb is the eigenfunction corresponding to A = 0, zero. One may calculate with [¢b[ = 1. If c10 is not zero, the implicit func— tion theorem gives a unique solution b(a) of (3.3) which depends on a analytically in a neighborhood of zero. b(a) is real for real a and b(O) = O. A more general analysis of (3.6) may be carried out by using Newton's polygonal method [ 1]. Consider a power series in two complex variables that converges in a neigh- borhood of (0,0). 2 _ ii G(a,b) — clOb-+cola-+c20b -+ -+cijb a + . Suppose the coefficients are real. We wish to find real solutions of G(a,b) = 0 corresponding to real a in a neighborhood of a = 0. Let ck0 be the first of the co- efficients c. 10 lemmas [ 1] apply: that do not vanish. The following two Lemma 3.7. Let k be even, Col #’O. Then if col/ck0 < O (col/ck0 > O), the equation G(a,b) = O has two different real roots b1(a), b2(a) for a > O (a < O) which are simple real roots, and has no real roots for a < O (a > O). 22 Lemma 3.8. Let k be odd and c01 # 0. Then G(a,b) = O has precisely one real root for both a > O and a < 0 and this root is simple. Under the above conditions, these solutions are given by fractional power series in ¥/a . The leading coeffients are given by ilf/Icol/Ckol f°r k even, and by k\""301;Cko for k odd. Since A1 ¥ 0, these lemmas apply to (3.6). The number and nature of the real roots b(a) will be determined by the first term ck0 #'O, the sign of Al/ck0 and the parity of k. we now conclude the proof of Theorem 3.1. (i) If no ci0 is nonzero, then (O,b) is a solu— tion of (3.3) for arbitrary b, and "vertical bifurcation" occurs. Therefore, we assume that C01‘ is the first such non- zero coefficient and apply the lemmas: (ii) If k is odd, n K’" I (3.4) bl(u) = k(/ -)‘17ck0 p + , where u and zl(a) = w(a,bl(u))-+b1(u) (111) Let k be even. If -Al/'ck0 > O r'bl((..L) = §/-A17Cko H + -°° , where H = +}(d (3.5) ( k k kb2(u) = ‘/-Al7cko H + --- , where H = - a . 23 If -Al/'ck0 < O, replace a by -a, and use Al/ck0 under the radical sign. In either case, 21(0)) = W(a,b1(u)) +bl(u) and 22(0) = W(a,b2(u)) +b2(u). In fact, the analytic function (3.6) z(()) = w(uk,b(u)) +bm) is a solution of (3.7) o = A(pk)z+N(pk,z) for all (complex) H in a neighborhood of zero. Here b(p) is any of the functions given by (3.4)-—(3.5) and fik = a. [In the case fik = -a, 'we replace H3 by -u¥ in (3.6)-—(3.7)]. For each a near zero, there correspond k distinct zi(a) (i = 1,...,k) corresponding to the k roots of a. These zi(a) form a cyclic system of solutions of the system (1.2). If a moves along some Jerdan curve about zero, the values zi(a) undergo a cyclic permutation when we return to the starting point. We also note that 2(p) may be expanded in a power series with real coefficents. 94. Stability In this section we give the proof of Theorem 3.2. Re- call the differential equation 24 (4.1) gi- = A(a)z+N(a,z) and the stationary problem (4.2) 0 = A(a)z-+N(a,z). We have shown that real nontrivial solutions 2(a) of (4.2) bifurcate from the trivial solution for real a near zero. These solutions may be found by substituting an appropriate real kth root of a (or -a) into an analytic function z(p) = Ylu-byzp2-+--- where Y1 is given in (3.4)-(3.5). For all p in a neighborhood of zero, 2(p) is an equilibrium solution of (4.3) %_zt_ = A(pk)z +N(p,k,z) if pk = a or (4.4) 3%;- = A(-uk)z +N(-uk,2) if (3‘ = -a. we wish to discuss the stability of the equilibrium solu- tions z(u) where u is a real kth root of a (or -a). Without loss of generality, we shall use p} = a in our calculations. Consider the linearization of (4.3) about z(p) for fixed p: (4.5) 3% = A(pk)y+D2N(pk,z(U))Y = L(u)y. For each p, L(p): Y 4 X is a continuous linear map. L(p) depends on H analytically, and LZHSy = L(fi)y. For H 25 sufficiently close to zero, L(u) may be regarded as a closed operator with dense domain D(A) in X; thus we may refer to Kato [11] for information about the dependence of its spectrum on u. Since L(O) = A(O), the spectrum of L(p) satisfies the following when H is sufficiently close to zero. (4.6) There is a simple eigenvalue A+(u) such that A+(O) = O and A+(p) is analytic in a neigh- borhood of zero. A+(p) is real for real p near zero . (4.7) The remainder of the spectrum satisfies [Rex] > 6/2 > 0. Our aim is to relate the sign A+ to that of A* when a is real and close to zero. Let y in Y be written as y = x-+§ where x E QY and g e PY z = x-tg. The eigenvalue problem for L(H) may be written and consider N(a,z) as N(a,x,§) where as the system r (4.8)(a) Ax [A(O) + (I -P)[A(p.k) -A(O) +DxN(pk,z(p))]x + (I-PO)[A(uk) -A(O) +D N(pk,z((_())]§ § (4.8)(b) A5 L + pmm“) -A(O) +D§N(uk,2(u))]§ P[A(u") - A(O) +DXN(uk, z (u) ) 1x or more simply 26 (4.9)(a) O [B(u)-l]X + C(u)§ (4.9) (b) 0 ll D(ulx + [EM -k]§ where B(u), C(u), D(p), and E(u) are linear maps for each u, and depend on H analytically in a neighborhood of the origin. Lemma 4.1. For A and p sufficiently small, B(p) and A-B(p) are invertible and 1 2 -1 n+1 +°°'+An[B([J.) ] +... (3(a) -)()'l = B(u)-1+A[B(u)- ] converges uniformly in the operator norm for A and p in a neighborhood of the origin. Proof. Since the set of linear homeomorphisms in L(Qy,0x) is open [ S] and B(O) = (I-P)A(O) is a linear homeomorphism, by proposition 2.1, B(p) and B(a)-—A are invertible for H and A sufficiently small. Also B(p) is close to A(O) in norm. The rest follows from the fact that if Tn are elements of Banach space and Q Q Zl)[lTnH < co, then ZETn converges.l’3 we may now solve (4.9)(a) for x in terms of g and substitute into (4.9)(b) to get 0 = [E(u) -D(u) (Em) - X)-1C(u) - x]: or 27 (4.10) 0 an.) -D(u)B(u)‘lcm) - A - D(u)[A[B(u)-1]2 + (A)2[B(u)-l]3 + - --]cm) f(u) -).->.9(u,>.) where f is an analytic function u in a neighborhood of zero, and g is an analytic function of H and A near the origin. A Newton's polygon argument shows that for real u there is a real solution A(p) = dluP-+d2p2p-+--- where d1 is the coefficient of the lowest power of p in f(p) = x(u) -D(u)B(u)-1C(u) and p is that lowest power. We now make the following observation. Recall that z(u) = w(px,b(u))-+b(u) satisfies 0 (I -P)[A((_(k)[w +b] +N(pk,w,b)] = Fl(u,w,b) (4.11) 0 ll P[A(uk)[w +b] +N(uk,w,b)] .-.- F2 (u,w,b). Let p near 0 be fixed. Differentiating F2(H,w,b) with respect to b at b(p) gives k k (4.12) P[A(u ) +D3N(p ,w(pk,b(u)),b(u)] + P[A(uk)%%+D2N(uk,w(uk,b(u)),b(u)]'g%(uk,b(u)). Since by I (3.6), 35% = _[DZF1(p,w(uk,b(u)),b(u)]-1[D3Fl(u.W(uk,b(u)bbhfll: 28 (4.12) is exactly E(p)-D(H)B(p)_1C(u). We have proved Lemma 4.12. For H sufficiently close to zero, we may determine the leading term of X+(H) by differentiating F2(p,w,b) with respect to b at b(u). The leading term of this derivative will be the leading term of A+(p). [Nete that F2(u,w,b) = O is the bifurcation equation.] Recall that the bifurcation equation may be written in the form _ 2 2 (4.13) F(a,b) — Alab+c02a +clob + where a = pk. Let cokbk+1 be the first term of the form ciob1+1 that does not vanish. In this case r k . . ./[A7cko| H + --- if k is even b(u)= < (k1/‘Vcko A if k is odd, uk=a. [For k even, pk = a if -A/ck0 > O and pk = -a if -A/'ck0 < 0.] Thus Ala+c02a2k +.-- + (k +1)cko[b(u)]k+ §-§(a,b(u)) . . k Ala-(k-+1)Ala-+ higher order terms in (A; -kAla-+higher order tenms in 5%; Thus d1 = -kA, p = k, and X+(H) is actually an analytic function of a: A+(a) = -kA1d-+--- . Recall that 29 1*(0) = Ala-+x2a2-+--- . Thus for real a near zero, A+ and A* have opposite signs. This completes the proof of Theorem 3.2. CHAPTER III A TWO-DIMENSIONAL NULL SPACE §l. Introduction Our aim in this chapter is to develop tools which will allow us to describe the set of nontrivial equilibrium so- lutions which bifurcate from the trivial solution of 3% = A(a,a)z +N(a,r3,z) at (0,0), where (a,8) is in C2. Part of the problem will be to choose suitable hypotheses so that the dependence of the bifurcating solutions on the parameters may be studied in a full neighborhood of (a,B) = (0,0). In the case which we shall study, A(0,0) has a two- dimensional null space, and A(a,8) has two eigenvalues A(a,B) and u(a,8) which pass through zero as (a,8) passes through the origin. The curves A(a,8) = O and “((1,8) = 0, defined for ((1,8) in R2, are assumed to cross transversally at the origin. This last assumption will allow us to change coordinates to T = A(a,B), n = “(0,5). we will formalize these hypotheses and discuss their implications at the end of this section. 30 31 In §2 we discuss the bifurcation equations and the coordinate change mentioned above. In §3 we give a simple example which demonstrates that the behavior of the bifur- cation set depends on the ratio Y = n/T (or T/nl- Moti- vated by this observation, we make a change of scale which allows us to introduce the ratio Y in place of one of the parameters ¢,n. Section 4-—6 discuss methods for analyzing the scaled equations and the interpretation of these results in the (¢,n)—space. ‘We now list the basic hypotheses of this chapter. (Hl)-(H3) are from chapter II and are repeated for conve- nience (H1) Let A be an open set in C2 that contains the origin and let X be a complex Banach space with dense sub- space D(A). Let A. ijD(A) 4.x be a closed linear operator with domain D(A) for each (a,B) in A, and let A be analytic in A in the sense that A(a,8)z has a Taylor expansion at each (ao,Bo) in A which converges in a disc [(b,a)-—(b0,ao)[ < r independent of 2. It follows from (H1), that if Y is the Banach space consisting of D(A) endowed with the graph norm, then for some neighborhood U of C2 we may regard A: U)(Y 4 X as a continuous linear map. (H2) Let N: U;(Y 4 X be a continuous map such that N(a,6,0) = O and D3N(a,B,O) = O for all (a,B) in U. 32 We assume that N is continuously Fréchet differentiable in a neighborhood V of (0,0,0). (H3) A and N are extensions of "real" operators in the sense that A(a,B)z = A(5,5)Z and N(a,8,z) = mass). In addition, we make the following additional assump- tions about A(a,8) = Aoz-+aAlz-+5Azz.l... (H7) Zero is an isolated eigenvalue of A0, and the generalized null space is two-dimensional. In addition, we assume that there are two distinct branches of eigen- values that depend on the parameters analytically and take on the value zero at (0,0). and “(0,6) H10a+polB+°°° are eigenvalues of A(a,8). Moreover, there are two dis- tinct eigenfunctions wb and Y corresponding to A(0,0) = O O = p(0,0), and we have the expansions m(a,8) = $b-leOa-tmblB-+ooo Y(a,B) = Yo'kyloan+y013.+... which converge in some neighborhood of (0,0). 33 (H8) For real values of the parameters, A and p are simple real eigenvalues (except when g(a,B) = A(G,B)), and ¥(a,B) = Y(a,3), $(a,B) = $(a,B). Furthermore, the curves A(a,B) = 0, u(a,3) = 0, defined for real a and B, cross transversely at the origin (see Fig. l); i.e., r- M ii 1 aa 38 det # O at (0,0) QB. he. aa 65 1dir Ahufiéffl) Figure l. Transversal crossing of zero eigencurves. This last assumption is important as it will allow us to make a local change of coordinates about the origin in JR2 in which the curves A(a,B) = O and p(0,5) = 0 be- come the axes. 34 Example 1.2. The hypothesis (H7) is necessary as the analytic dependence of the eigenvalues and eigenfunc- tions upon the parameters does not necessarily follow from that of A. Let L_B -a map C2 into C2. The eigenvalues are A = a2-+82 and H = -./ 2-+82 which are not differentiable at (a,B) = (0,0). The problem is that the eigenvalues are given by different branches of the same multivalued function a2-+82 and (0,0) is the branch point. §2. The Bifurcation Equations Let P be the projection given by proposition II 2.1, and PX’ Py, QX’ Qy. be the subspaces described in §2 of chapter II. In our case Px = PY = span[qo,Yo} and is homeomorphic to C2. The projection P: X 4 PX is the sum of two projections, P1 and P2. P1: X 4 Span{¢o} = P; 2 1 2 and P2: X.4 span[YO} = Px where Px = Px:® PX‘ If 2 is in X, we may rewrite 2 as 2 = w-+x-+y where w 6 0x, 2 x 6 P; and y e PX’ As in §2 of chapter II, the equation (1.3) is equivalent to the system 35 ”(2.1) (a) o Aow + (I -P)[Ao-A(a,f3)] (w+x +y) 'l" (I " P)N(G,B,X,Y,W) (2.1)(b) 0 = Pl[aAl-+BA2-+-°-](w-+x-+y)-+P1N(a,B,x,y,w) [(2.1) (c) o = P2[aAl+BA2 +---](w+X+y) +P1N(a,f3,x,y,w) Furthermore, we have shown in chapter II, §2, that there is a unique w(a,8,x,y) such that w(0,0,0,0) = O and (a,8,x,y,w(a,B,x,Y)) is a solution of (2.1)(a) for (a,B,x,y) near (0,0,0,0); w(a,8,x,y) can be expanded in a power series in some neighborhood of the origin. Also §(a,f3.x,y) = “5.5.5350. As before, we may show that g(o,o,o,0) = %‘§(o,o,o,0) = 0. Moreover, since w(a,B,0,0) = 0 for all values of (a,fi) near (0,0), the uniqueness of w implies that this power series will contain no terms which contain only a or 8. Thus w(a,B,x,Y) is of the form W(G,5,X,Y) = Clax-fczfix-+c3ay-+c48y-+c5x2 + cexy-+c7y2-+higher order terms in x,y,a,B. we may substitute this expression for w into the bi- furcation equations (2.1)(b) and (2.1)(c), and obtain two power series in 4 variables: (2 2) o P1[aAl-+BA2]x +-Pl[dAl-+BA2]y-+-.- O P2[aA1-+BA2]x + P2[aA1-+8A2]y-+-o- . 36 We now determine the terms listed in (2.2). Differ— entiating the expression 0 = P1[A(a,f3) -A(a,B)]cp(a,t3) . - .QA _ With respect to a at (0,0) gives 80(O’O)¢0 — PlAlmO. A similar argument shows that $3“), 0)ch = P1A2°~°o 31*. = 3E. _ a“(dome P224140 and 63(0’0)Y0 P2112110 Also P2A1cqO = PzAchO = PlAlYO = PIAZYO = 0. Thus the bi- furcation equations become ( o = abs-g}: +B%%X+P1[aAl+fiAl]w + Pl[A(a,B) —A0 -aAl -BA2][w+x+y] +P1N(c,8,x,y,w) (2.3)( H o = a%§X+B§-5x+P2[aA1+BAl]w t + P2[A(a,8)-Ao-aAl-8A2][w-+x-+y]-+P2N(a,8,x,y,w) Henceforth, we will seek real solutions (x,y) of (2.3) corresponding to real parameters (a,B). we may as well assume that the norm is Euclidean. It is convenient to make a change of variables in the parameter plane so that the curves A(a,8) = O and u(a,8)==0 become the axes. Define the transformation T: (a,B) 4 (T,n) by T = A(336): T] = “(aaB)o Since A(a;B) = 0 and “(0,5) =0 37 cross transversely at (0,0), T is a homeomorphism in a neighborhood of (0,0). Then ... PM m'fi—lr- 1 a.) ad OB T 2 = +0!(T,n)| . 314 A! [34 _aa 36‘ [TL Thus terms which are first order in a and B, are first order in T and U7 terms which are second order in a and B, are second order in T and n; etc. The bifurca— tion equations take the form 0 Tx 2nd order terms in x and y + higher (2 4) + order terms in x,y,T,n O = ny There are no terms which contain only powers of T and n. Terms which contain only powers of x and y are de- rived from the nonlinear term which we refer to as N(T, ngx3Y:W(x:Y: Tafl))° §3. An Appropriate Change of Scale As motivation for what we are about to do, consider the following. Example 3.1. Consider the system O = TX + %y2+x2 (3.1) l 2 2 0 = ny +-Z>c +37 38 The numbers of nontrivial solutions of (3.1) corresponding to each (T.n) are given schematically in Figure 2. l in l \\\‘ l ‘A ‘7’ 3 Figure 2. Dependence on (T,n) of the number of nontrivial solutions of (3.1). As we can see, the number of nontrivial solutions depends on the ratio T/n or (n/%). Also note that it is impos- sible to apply the implicit function theorem to obtain solutions of (3.1) of the form (x(T2n):Y(T:n):T:n) in a neighborhood of (x,y,¢,n) = (0,0,0,0). New consider the following change of scale. Let T = Yn, and replace x by fix and y by ny in (3.1). After division by n2, (3.1) becomes 39 r O = YX +%Y2+X2 02) ( 2 y.%a+y L.O Sectors in the (T,n)-plane and the corresponding re- gions in the (Tby)-plane are represented in Figure 3. Points on the Y-axis correspond to different slopes through the origin in the (T,fi)-plane. // 43$ 47 - ,,__.:\\ Rs 45/ / 4‘ $4, // 4 Figure 3. Correspondence between (T.n)-plane 7 // and (n,Y)-plane. The numbers of nontrivial solutions of (3.2) correspond- ing to each (U,Y) are given in Figure 4. 4O :////////W//’/4 ///////3////A Figure 4. Dependence on (n,Y) of the number of nontrivial solutions of (3.2). It is easy to check that at most of the solutions (XO’YO’YO) of (3.2), the implicit function theorem applies. The points at which it fails correspond to the lines in Figure 4 which indicate a change in the number of solutions. In what follows we shall give hypotheses under which a similar scaling may be determined for the bifurcation equations (2.4). Solutions (x,y) of (2.4) which tend to (0,0) with (T:n) will correspond to solutions of the scaled equation which remain bounded as (7,n) approaches (0,0). In order to accomplish this, we shall require the bifurcation equations to be of a certain form so that we 41 can obtain an a priori estimate of the magnitude of solu- tions of (2.4) which tend to (0,0) with (T,n). This a priori estimate will determine the scaling we choose. Let T = n = O, and suppose that the first terms of (2.4) in x and y which do not vanish have the same de- gree q in each equation. The terms of degree q form a continuous q-linear form Mq: 1R2q4 ZIR2 ' n n-1 n-1 nf (3.3) aox -+a1x y-+--w+an_1xy -+any Mq((x,y),..., (x,y)) = n n-1 n-1 n box -+blx y +-.-+bn_lxy -+bny‘d we say that Mq is nondegenerate if (3-4) Mq((X:Y):-°-: (XJY)) # (0:0) for all (x,y) 7! (0,0) in 122. Since the unit ball in R2 is compact and Mq is contin— uous, (3.4) implies there are positive constants c1 and c2 such that (3.5) all (x,y))q _<. Mq((x,y),..., (x,y)) 3 c2 [(x,y) (‘1 for (mil) in R2. (H9) We assume the bifurcation equation has the form (3.6) 0 7x +S(T:T]:X:y) +M ((X,Y),..., (X,Y)) +]R(T:T]:X3Y) O 111/ q 42 where S(0,0,x,y) s O, S(T,n,x,Y) contains terms in x and y of order less than q and there is a positive con- stant c3 such that OJ) [Nnm&W\g%HnmLH&W| when (T,n,x,Y) is sufficiently close to (0,0,0,0). We also assume that IR(Tk: le: xk’ Yk)l (3.8) 4 0 for all sequences Hampfl [7k]: [nk}, [Xk}, {Yk} that tend to zero with [(xk,yk)[ #‘o for all k, and that Mq is a nondegenerate q-linear form. Proposition 3.1. Let (H9) be satisfied. Then for any (T,n,x,y) sufficiently close to (0,0,0,0), such that (x,y) is a solution of (3.6) corresponding to (Tan) # (0,0), we have Hmw(ngVWRJT where m is a constant independent of ¢,n,x,y. Proof. Suppose not. Then there exist sequences [Tk], ink}: {Xk}, [yk] that converge to zero, with (Xk’yk) #‘o, (7k3nk) # (0,0) and Hflwflfld 1(Tk, nk)‘ >k. 43 We may assume that Tk’ nk’ xk, and yk are small enough so that the estimates (3.5) and (3.7) hold. Consider 1 Tkxk | (xk, yk) lq nkyk O + S(Tk: fik’xk’yk) Mq((xk:yk):---: ()ck’yk» R(Tk’nk’xk’yk) + + ‘(x-k: Yk) ‘q )Xk’ yk) ‘q As (Tk’nk) and (xk,yk) approach the origin, so does the last term. Estimate (3.5) implies that there are positive constants m1 and m2 such that |(xk,yk) (‘1 ml‘g S.m2 Tkxk leYk + S(Tk’ 'flk: xk’ yk) for k sufficiently large. Thus by (3.7), m 2 [(xkdknq 2 )(TRXJQ, myk)‘ +c3lTk: TR) " ‘ (xk: YR” ((kak) (‘1 Z (1 +C3) ‘ (Tk: 71k) ‘° ‘ (xk’YkH and -l [(inyk) [‘1 ‘1 +c3"”“3 2 wkmkn which is a contradictioan 44 Using the a priori estimate of proposition (3.1), we will now exhibit appropriate scalings of the system (3.6) for different sectors of the (¢,n)-plane near (0,0). Let 9p = UT,TQ: [n] > plwl} for p > 0. Let (x,y) be a solution of (3.6) corresponding to (T,n), that tends to (0,0) with (T,n). Then x and -——JL-- q‘lff‘n) q‘IJ—I) n remain bounded as long as (T,n) remains in Qp. Thus none of the bifurcating solutions corresponding to (T:fi) in Qp are lost if the following changes of scale are made. Case 1. If q is even, let n = gq'l, T = ygq'l for |y| g %- and replace x and y by gx and gy respec- tively. Then (3.6) becomes 0 yx (3.9)(a) = + Mq((x,y),...,(x,y))-+§§(x,y,Y,g) 0 Y where R is analytic in (x,y,Y,§). [The transformation of the sector Op in the parameter plane under this change of scale was given in Figure 3]. Case 2. q is odd. In this case, we must distinguish between D > 0 and U < O. 45 To study solutions corresponding to n > 0, we use the same change of scale as in case 1. Here, however, we must keep in mind that each U corresponds to j;§, and there will be an "extra set of solutions" for g g 0. To study solutions corresponding to U < 0, we let n=-§q-1, 'r = -Y§q-l gy for |y[ g %=. Then (3.6) becomes , and replace x by gx, and y by Y O x = (3.9)(b) 0 = - M (0,0,(x,y),---,(x,y)%+§R(X,Y,Y,§) O l y q where R is analytic in (x,y,y,§). Again, we will disregard the extra branch of solutions corresponding to g < 0. To study solutions of (3.6) for (T,n) near the T-axes, the roles of T and n are reversed. In all cases, we must now determine the behavior of real solutions (X,y) corresponding to real (y,§) for systems of the form -1 YX+aan+an_1Xn 17+". +aOYn+§Nl(X,Y,Y, g) 0 (3.10) n n-l n y-+bnx -+bn_lx y-+----+boy -+§N2(x,y,Y,§). 0 ll §4. Solutions of Type I - Fold curves Consider the system (4.1) n n-l n O = F(x,y,y,§)==yx-+anx -+an_lx y-+--o-+a0y -+gNl(x,y,y,§) n n-l n o = G(x,y,Y.§)==y-+bnx -+bn_lx y +°°°-+boy -+§N2(x,y,v,§) 46 where Mn((x,y),...,(x,y)) = n n—l n b bnx +bn_1x y +--- +b0y .1 is a nondegenerate continuous n-linear form, Mn: It 4 It; and N1 and N2 are analytic in (x,y,y,§). Let F F x y (4.2) J(X,y,v,§) = det at (x,Y.Y,§). GX Gy For g = O, (4.1) and (4.2) become n n-l n O = P(x,y,y) = yx-ranx -+an_1x y-+----+aoy (4.3) n n-l n 0 = Q(x,y,y) = y-kbnx -+bn_1x y-+-o--+boy and (4-4) 0 = J(X,Y,Y,O) = PXQy-QXPY at (X:Y:Y:o)- Let (XO’YO’YO) be a solution of (4.3). If J(xo,yb,Yo,O)7¥ 0, the implicit function theorem guarantees a unique solution (x(y,§),y(y,§),y,§) of (4.1) in a neighborhood of (yo,0) such that x(y0,0) = x0, and y(yo,0) = yo. J(x,y,y,§) #’O for (x,y,y,§) near (xo,yo,y0,0). Thus the solution (XO’YO’YO’O) determines the behavior of nearby solutions of (4.3). 47 In the event that J(xo,yo,yo,0) = O, we wish to impose conditions on (4.1) and (4.2) so that we may still determine the local behavior of solutions near (xo,yo,yo,0): (H10) We assume that for all (XO’YO’YO) satisfying (4.3) with (x0,yo,y0) # (0,0,0), we have rank = 2 at (xo,yo,yo,0). (H10) guarantees that about each nontrivial solution (xo,yo,yo,0), there is a neighborhood of solutions to (4.1) that is homeomorphic to an open set in IR2 . [The point (0,0,0,0) is a special case which corresponds to studying solutions which bifurcate from the trivial solution near the T or U axes. This case is banished to section §6.] (H11) we assume that for all (XO’YO’YO) which satisfy both (4.3) and (4.4), the matrix has rank three at (xo,yo,y0,0). A solution (x0,yo,y0,0) of (4.1) and (4.2) for which either 48 F F F X y Y 4.5 d t G G G o ( )(a) e x y Y 3’ J J L X Y Y. or F F F x y *3 4.5 b d t G G O ( H) e ex y g 9‘ J J J -x y 5, will be said to be of type I. When (4.5)(a) holds, these solutions have the pleasant feature that local behavior of nearby equations is completely determined by the polynomials (4.3) and (4.5), since (4.5)(a) is exactly the condition ’ '1 Px PY }{ (4.6) det Qx Qy O 7’ O J L.x Y QY‘ Note that (4.5)(a) implies Gx GY det #'O and x #’O. Jx Jy Theorem 4.7. If (XO’YO’YO) is a solution of (4.3) and (4,4),and (4.6) holds, then there is a curve y(§) and a solution curve (x(§),y(§)) such that y(O) (x(O),Y(0)) = (X = YO, O,yo). When y crosses y(§), the number 49 of solutions of (4.1) near (x(g),y(§)) changes by two. we will refer to y(g) as a fold curve. If (4.5)(a) fails and (4.5)(b) holds, we may prove a similar result. In this case the solution (x O) 0: YO: YO: of (4.1) determines curves (x(y),y(y)), g(y) with (x(yo),y(yo)) = (XO’YO) and g(yo) = 0. In the following we shall concentrate our attention on (4.5)(a), as results corresponding to (4.5)(b) are similar. We isolate a special case in the following lemma. Lemma 4.8. If (XO’YO’YO) is of type I, the number of solutions (x,y,y) of (4.3) near (XO’YO’YO) changes from zero to two as y passes through YO' Proof. Without loss of generality, we will assume that det y #’0. Thus PY and Qy are nonzero at (xo,y0,y0), and we may use the implicit function theorem to guarantee the existence of y(x), y(x) for x near x0, is a solution of (4.3) and y(xo) = y0 and y(xo) = such that (x,y(x),y(x)) Yo' we shall show that y may be written in the form y(x) = Y0+a2(x-xo)2 +O(x-x0)3 ..l where a2 — 2-y”(x0) 50 (i) Computation of y’(xo). By the chain rule, "' I 1 r 1 ’- fi 1! (x0) QY —PY Px = l ’ PYQY - PYQY Ly (XOL _-QY Py __ L- Qx r- m QYPX -— Pny = 1 P - P YQY YQY P P - + t QY x YQX_. I Qny — Pny - 1 at x _ P _ , YQY Pyoy O o t .1 I = Thus y (x0) 0. (ii) Computation of y”(xo). Further implicit dif- ferentiation plus part (i) gives P +2P 91+P QXZ+P d—2-X+P d—2-I--O at (x ) xx xy dX yy(dx y dxz y de _ O’YO’YO ' dx2 y yx if and only if 2 d 1 2 Also __Y.= Q- [Qxx-ZQ (gfi + Q Q2) ]. Thus y”(xo) = 0 (4.9) o = Qy[Pxx +2ny _ 92 Py[Qxx-+ZQ + Q 51 By (4.6), O # Jny--JYQx Q = [Pxey'H‘Dnyx"Pquxuprxx:l y -[p +PQ -PQ-PQ]QX nyy x YY YY x Y xY 2 =P +2 P -2P — P xey Qxy ny nyny Qxx yQy 2 P Q + P Q2 - Q _Y_§. YY X YY Qy (2x (2x 2 QY{QY[PXX ' ”mu-2;) + PYYKQ) ] (2x ox 2 ' Pymxx ' 201675;) + (INK-6;) J} 2 BY; .22 QY[QY[Pxx +2ny dx + PYYKOX) ] 2 +°yy\%) 1}. filfi‘ - pymxx + ZQxy Thus y”(x0) #'O.[] Proof of Theorem 4.7. Once again, we assume P x det Y #’O. Q 0 Since (4.6) holds, there is a unique curve (x(g),y(§),y(§),§) of solutions of (4.1) and (4.2) with x(0) = x0, y(0) = yo, Y(0) = YO' For E sufficiently close to zero, (4.6) still holds. Moreover, 52 det Y Y = O at (x(E),y(§),y(E),§). Fix E and consider the functions r-§(x.vY.v'Y) = F(X:Y:Y:§) (4.10) ' 4 6(X3Y3Y) = G(XJYJYJE) L3(X:Y:Y) = J(X,Y,Y,E) (4.10) satisfies r' 1 'UI "Ul WI (4.11) det 0| (0| 0| x y Y ‘r’0 at (x(E),y(E),y(E)). cll c-ll q: L" Y Y.) ‘we now do some calculations similar to those in Lemma 4.8 and use (4.11) in place of Jka-Jny #’O to reach the desired conclusion. Thus for fixed E close to zero, y(x,§) has the form - — — 2 — 3 ((095) = y(g) +a2(x-X(§)) +0(x-X(§)) where a2 depends continuously on E and hence has con- stant sign for g near 0. Thus there are two solutions of (4.1) near (x(g),y(§)) on one side of y(g) and no nearby solutions on the other.C) 53 We have the following local picture when (y,§) is near (yo,0) [Figures 5 and 6]. 4A “1v Figure 5. Behavior near a fold curve in the (g,y)-plane. Figure 6. Behavior near a fold curve in the (T3n)-plane. 54 These pictures are local representations. It is en- tirely possible to have other solutions (xl,y1) that correspond to YO' However, the solutions near (x1,yl) will correspond to a different branch which does not in- tersect the branch determined by (x0,y0) in some neigh- borhood of (y0,0). §5. Solutions of Type II Let (H10)-(Hll) hold. Solutions (XO’YO’YO’O) of (4.1) which satisfy neither (4.5)(a) nor (4.5)(b) will be called solutions of type II. In general, the local behavior near solutions of type II is not as easy to determine as that near a solution of type I. waever, we can give a complete description of this behavior when the original bi— furcation equations contain a nondegenerate bilinear form when T = n = 0. We shall determine some general properties of solutions of type II. If (XO’YO’YO’O) is a solution of type II, we must have Px Py (5.1) det Q Q = 0 at (xo,yo,vo). X Y Proposition 5.1. If Px = Py = 0x = QY = 0 at a solu— tion (XO’YO’YO) of (4.3), then either (XO’YO’YO) (0,0,0) or the n-linear form is degenerate. 55 Proof. At (xo,yo,y0), we have _ _ n n-l ... n O - xOPx-tyOPY - yoxo-+n[anxO-+an_lxo yO-t -+aoyo] n-l n n +bn-lx0 YO + - . - +boy0] O = XOQx-tyOQ = yO-+n[bnxO Y Since P(xo,y0,yo) = Q(xo,yo,yo) = O, we have (n--l)yO = 0 and (n--l)yoxO = 0. Thus yo = 0. If Y0 = O and x0 # 0, then an = bn = O and a degeneracy occurs in the n-linear form. If x0 = 0, PX = 0 implies Y0 = 0.C) Recall that (0,0,0,0) is not considered in (H10)- (H11), and therefore cannot be a solution of type II. Be- fore we go on, we shall consider two special cases. Case 1. x0 = 0. If a0 #'0 in (4.1), then y = O and J(0,0,yo,0) = YO cannot be zero unless Y0 = 0. If a0 = 0, then .1. _ n-l YO - (-1A0) o [If b0 is also zero, the n-linear form M.n is degenerate]. Suppose yO is real. Then the matrix in (H11) becomes ' W YO -al/b0 O 0 N1 (031(0) Y030) (5.2) «bl/bO -(n-l) 0 N2(O,yo,yo,0) J J - n - 1 J _ x Y ( ) g A Now J(0,yo,yo,0) = 0 if and only if Y0 = al/bo. Case 1 can only occur if a0 = O and b0 # O. In this case 56 (H10)-(Hll) are satisfied if and only if Nl(O’YO’Y0’O) # 0. Then r- 7 det (1 ’n) 0 N2 (0:170: Y020) 7! O J (l-n) J L Y s - and W O N1(O’YO’ Y030) i det 1 #VO. (1 ‘11) N2 (0: Yo: YO’O)_) Case 2. ox = Qy = 0 at (x ) 7‘ (0,0,0). As in o’yo’ *0 the proof of proposition 5.1, we have that y0 = 0. If bn #'0, then x = O and J(0,0,yo,0) = YO' Thus as before, we assume bn = O and an #'0. Then (XO’YO’YO) # (0,0,0) implies __L_ _ n-l . xO — (-YO/en) Wlth Y0 #’O. The matrix of (H11) becomes (5.3) F ] Yo‘nYo -an-1YO/ an xo N1("o’ 0’ Yo’ 0) O l—Yobn-l/an O N2 (X0, 0, YO’O) 2 n-2 [:Y0(n-l) bn-le Jy l-YObn—l/an JE J QY = 0 implies Y0 = an n-1 and bn-l #’0. Thus Case 2 occurs only when bn = 0, an #’O and bn-l #’O. In order 57 for (H10)-—(Hll) to hold, we must have N2(xo,0,yo,0) #'0, x0 #'0, and Y0 # 0. Example 5.1. Case 2 may be reduced to Case 1 by a change of variables. we shall demonstrate this in the case where M.n is a nondegenerate bilinear form. If we recall the form of the nonlinear term before the scaling of §3, we may write 2 2 o = YX+a2x +alxy+aoy +N1(XJY)Y§3§) (5.4) 2 O = y +blxy +boy +N2 (X:Y:Y§:§) Then (5.5) x = _ fiL Y = aZ/bl’ y = O is a solution. 1 Now consider pu-Tbouz-+blvu-+Nl(v,u,o,po) 0 II (5.6) O = v+aov2 +a1uv+a2u2 +N2 (v,u,o,p0) If (xo’Yo’Yo’go) is a solution of (5.4) with y #'0, then u = y/&, v = x/y, p = l/y, o = y§ is a solution of (5.6). The solution (5.7) becomes 11 = 0, v = -l/a2, p = bl/aZ’ This correspondence arises from interchanging the roles of T and n in the scalings discussed in § 3. of 58 Proposition 5.2. Let (x 0) be a solution 03 Yo) Yo) type II with x0 #'0. If 0x = O and QY #'0, then Px = J2 = 0. If 0x #'0 and Qy = 0, then PY = Jy = 0. Proof. PXQy--QXPy = 0 and (4.5)(a) fails if and only Since (4.5)(a) and (4.5)(b) both fail for a solution of type II, we have Px x F§ (5.7) det Qx o G); a" o J J J t x Y 5.) or PP x F '1 Y 5 5.8 det O Gr 0 ( ) Qy g 7‘ J J J _ Y Y 5.. Proposition 5.4. Let (XO’YO’YO’O) be a solution of type II. Then one of the following occurs. (a) (5.7) holds and' Px F PX x det § # 0 or det #’0 OX G5 (.0X 0 (b) (5.8) holds and P F P x det Y 5 a! 0 or det Y 7! 0 Q G Q 0 Y E Y 59 Proof. This follows from Case 1, Case 2 and proposi- tion (5.2).:) Definition 5.1. Let T: R24 R2 be given by (u,v) 4 (f(u,v),v). We say that (uo,vo) is a cusp point for T if i: if. (i) Bu(u0,v0) = 0 and all2(u0,v0) = 0 2 .. a f (ii) auav(uo’v0) #’O (iii) §3§(uo,vo) #'0. au Proposition 5.4. Let (x0,yo,y0,0) be a solution of type II. Then (a) The implicit function theorem enables us to solve (4.1) for at least one of the following in a neighborhood of (XO’YO’YO’O): y(x,§), y(Y,§), g(x,y), gum). (b) At least one of the following maps exists and also satisfies (i) and (ii) of definition (5.1) at (XO’YO’YO’O) (x,§) 4 (y(x,§),§) (ing) 4 (y(y,§),§) (x,y) 4 (§(X,y),y) (Y.y) 4 (g(y,y),y). Proof. We will only give the proof in the case that Py X Fg (5.9) det Qy 0 Gg 7! 0 J,Y Jg 6O Py x and det # O at (x Qy 0 031,0: YoJO) . All other cases are similar. Since XOQy #’O, there exist unique real analytic func- tions y(x,§) and y(x,g) such that (x,y(x,§), y(x,§),§) satisfies (4.1) near (x0,0) and y(x0,0) = O, y(xo,0) = 0. Recall that x0 #’0 implies Qny-Qka = O for a solution of type II. As in Lemma 4.8, we have (a) fl(x0,0) = O and %§(x0,0) = -Q /Q ax X Y 3.3! (b) 8x2(xO,O) = 0 if and only if Qny-Qny = O 3.1 31 (C) sz (x ’0) = 3y aa+ Jy at+ J): axag 0 -XOQY _ Jy[xOG§] -Jv[PyG§ -:§Qy] +JE[xOQy] . (xOQy) 2 C ' Thus (5.9) implies g§§%(xo,0) #'O.D Definition 5.2. If (XO’YO’YO’O) is a solution of type I, we will call (x 0) a fold point. If 0) Yo) Yo) (x O) is a solution of type II, and all conditions 0: YO: Yo: of definition (5.1) are satisfied for one of the maps guar- anteed by proposition 5.4, we will call (x 0.0) a o’YOJV cusp point. 61 In general, we shall not be able to verify the third condition of definition (5.1), as it depends on the order of contact between the two curves P(x,y,y0) = O and Q(x,y,yo) = 0 at (x0,yo). On the intuition that two conics cannot "touch too much“ without some form of degen- eracy, we now specialize to the case of a nondegenerate bilinear form Ax2-+Bxy-+Cy M2[ (x,y), (x,y)] = axz-tbxy-Tcy where A, B, and C are not all zero. Consider the system, 0 F(x,y,y,§) Yx+Ax2+BXY+CY2+§N1(X:Y:Y:§) (5.10) 0 = G(x,y,v,§) y+ax2 +bXY+CY2 +§N2 (3910165) As an example of what to expect, we will look at Case 1. In this case C = O, c #’0 and (XO’YO’YO) = (O,-l/c,B/c). The matrix (5.2) becomes 1 [ o o 0 N1 (0,-1/c,B/c,0) -b/c -1 0 N2 (0,-1/c,B/c,0) L-2A + Bb/c -B -1 J: 3 For a type II solution, we must have -A-+Bb/c = 0. 62 By the implicit function we may solve (5.10) for g(x,y) and y(x,y). If we eXpand g(x,y), y(x,y) and the equations in (5.10) in power series and then equate coefficients, we find that 31 — _ (5.11) ax‘XO’Yo) - Qx/QY 331 -1 41 31 2 (5'12) ax2(xO’YO) = -Qy [Qxx-+Qxy ax + QYYKBX) ] _ 2 21 ex 2 ‘—- Qy[a-+b ax + CKBX) ] 2 (5.13) 6x2(xo,yo) = 0 if and only if Px-Pny/QY = 0 2 (5.14) 5—125-(x0,y0) = o if and only if 8X 2 ° = [Paws £26 + 54%)] 2 _ '1 .31 31 PYQY [Qxx'l'Qxy BX + QYYKBX) ] 2 31 31 _ - gr 2[2r>.+sax + CKax) ] 2PyQy1[a +b x+c[§1x)2] 3 (5.15) 'i-§(x ,y ) = 0 if and only if 3 0 O ax -Qx 52 -Q 2 = _ _.1_ _1 + J El— 0 [ny nyK :flax PyQy [Qxy QYYKQY flaxz 2 2 if and only if, J = 0 or h—2 = 0. However, Q—x(x ,y )== y 2 2 0 0 Bx 8X 0 implies that (x,y) = (l,%§) is a nontrivial solution of ‘I‘l'll 63 O = axz-I-bxy-i-cy2 O = sz-i-Bxy-i-Cy2 Since the bilinear form M2 is nondegenerate, this implies Jy = 0. Since 0y #'0, we must have JX = O. This gives the following -B=O -2An+B(b/c) = 0 and -An+Bb/c = 0 Hence A = O = B.D Because of example 5.1, we may neglect Case 2. It remains to examine the situation where x0 #'0, and one of 0x or Qy is nonzero. By proposition 5.3, one of the following must hold r- a-n Py X F§ Py x (5.16) det Qy O Gg 710 and det 7’0 Q 0 J J J L y Y 5. y or i- -1 PX x Fg Px x (5.17) det Qx C) Gg # O and det #'0 Q 0 J J J x t x v t, at (x0) Y0) YO, 0) 0 64 we shall only consider (5.16), as essentially the same result holds for (5.17). We use the implicit function the— orem to obtain y(x,§) and y(x,§) such that (x,y(x,§), y(x,§),§) is a solution of (5.10) in a neighborhood of (x0,y0,y0,0), and y (x0,0) = yo, y(xo,0) = YO' The analogues 3 of (5.11) - (5.15) now hold, and if L40: 5X yx-tsz-thy-+Cy2 0,0) = 0, the system 0 = P(x,y,v) (5.18) y+ax2+bxy+cy2 0 = Q(X:Y:Y) must satisfy J(xo,yo,yo) = PxQ --QXPy = 0, Jx = J = 0 at Y Y (x0,yo,yo). ‘we will take care of this situation with the following lemma. Lemma 5.5. Given A, B, C not all zero, there does not exist (x,y,y) # (0,0,0), x #'0 and a, b, c such that M2 is nondegenerate and the system in (5.18) satisfies )==J =J =0. Proof. we must satisfy the equations (1) o = YX+AX2+BXY+CYZ .. 2 2 (ii) 0 = y-Tax -+bxy-+cy (iii) 0 = J(X,y,Y) = y-fybx-tZycy-tZAx-tZsz + 4Acxy+By+2ch2--2an2--4any--2bCy2 (iv) 0 = J x yb-TZA-t4Abx-+4Acy-4an-4aCy (V) o = Jy 2yc-T4Acx-PB-+4ch-4aCx-4bCy 65 We may use the fact that xe +yJY = 0 to replace (iii) bY (vi) 0 = y - 2Abx2 - 4Acxy - 2ch2 + 2an2 + 4any + 2bCy2 . Since x 7’ O, we may multiply equations (iv), (v) and (vi) by x and use (i) to eliminate Y. We now have four equa- tions which are linear in a, b, and c, F O = ax2 +bxy+cy2 +y 3 3 0 = a(2Bx +4Cx2y) +b(-2Ax +2Cy2x) +c(4Ax2y -ZBy2x) J + (-Ax2 - Bxy - Cyz) O = a(-4Bx2 - 4ny) +b(-Ax2 -Bxy -Cy2 +4Ax2) +c(4Axy) +2Ax k 0 = a(-4Cx2) +b(-4ny) +c(2Ax2 +2Bxy-2Cy2) +Bx. This system may be reduced by the usual row operations to the following [o 0 ax2 +bxy + cy2 +y 3 2 b(Ax +Bx y+ny2) +Ax2 +Bxy-Cy2 3 2 0=b(3Ax +3Bx 2 (5. 19) 1 y + 3ny2) + c (4Ax y + 4By2x +4Cy3) + 2Ax2 + 4Byx + 4Cy2 k0 c (2Ax2y + ZBxy + 2Cy2) + Bx + 40y. An appropriate linear combination of the last three equations shows tha t 66 3[Ax2-+Bxy-Cy2]-+2y[Bx-+4Cy] - [2Ax2 +4Bxy +4Cy2] 0 ll sz-TBxy-TCyZ. Now (5.19) implies O = Ax2+Bxy--Cy2 0 = sz +2Bxy +2Cy2 0 ll Bx-T4Cy. Thus Cy = O = Bx = Ax. Since x y'o, we have A = B = 0. If C #’O, we must have y = 0. Then (ii) implies a = 0 and CY M2[ (x,y) (x,y)] = bxy+cy2 has the nontrivial solution (x,0), and therefore is degen- erate. we have just proved the following: Theorem 5.6. If the bilinear form Ax2-+Bxy-(-Cy2 M2[(x,y), (x,y)) = ax2+bxy+cy2 is nondegenerate and A, B, and C are not all zero, then all solutions (XO’YO’YO’O) #’(0,0,0,0) of 67 F0 = yx+Ax2 +BXY+CY2 +§N1(X’Y’Y’ g) (o [_0 y-Tax2-+bxy-+cy2-+§N2(x,y,y,g) J(XJY: Y: S) are either cusp points or fold points, provided that (H10)- (H11) are satisfied. we shall use the following example to demonstrate be- havior near a cu3p point. Example 5.2. Consider the system (5020) (a) O YX+XY+TN1(X:Y:Y:T) (5.20)(b) o 2 2 y+x +y +TN2(X:Y:Y:T) where N1(O,-l,l,0) #'0 At the solution (O,—l,l,O), the matrix corresponding to (5.2) is given by O -1 0 N2 (0,-1, 1,0) O -1 -1 J L 5 3 and has rank three. By theorem 5.6, (0,—1,l,0) is a cusp point. If we solve (5.16)(b) for y(x,y,T), we obtain y(x,y,T) = -1-+x2-+higher order terms. substitution into (5.20)(a) gives (5.21) o = (Y-1)X+X3+TN1(X:Y(X,Y,T);Y:T)- 68 we make the change of variables § = -(y-1) T = -TN1(x,y(x,y,T),y,T) in a neighborhood of (O,-l,l,0). Since Nl(O,-l,l,0) is nonzero, this change of variables is a homeomorphism by the inverse mapping theorem. Now (5.21) becomes The mapping T: R 4 1R2 given by >4 :4 l .< N 41 (5.22) 4 <1 <1 <1 is the standard form for the cusp singularity [ 3]. we have the following local picture, [Figure 7] ‘N T\ Figure 7. Behavior near a cusp. 69 The number of real solutions (x,y) corresponding to each (Y,T) in a neighborhood of (1,0) is given schema- tically by the following bifurcation diagram [Figure 8]: Figure 8. Bifurcation diagram in the (T,Y)-plane. If we change back to the (T,n)-plane by setting U = YT: the corresponding diagram is given by Figure 9. The curve r in Figure 8 has become a curve f tangent to the line n=T- A q 4 Figure 9. Bifurcation diagram in the (T,n)-plane. 70 Note that this is only a local picture, and is valid only in a neighborhood of the curve T. The picture can be completed only when we have analyzed the behavior of solutions near every (x which is a solution 0’ Yo, YO: To) of 2 2 {0==yx+yx O=y+x +y . In general, if T: R24 1R2 satisfies definition (5.1), there are local changes of coordinates in the domain and in the range of T such that the local picture is given by Figure 7 and T has the form (5.22) [ 3]. The hypotheses (H10)-(Hll) were chosen to insure that the set of solutions (x,y,y,T) of (4.3) is a two manifold in the neighborhood of (xo,y0,y0,0) #'(0,0,0,0). That is, there is only one solution branch passing through (XO,YO:YO, always happen in applications. If two solution branches pass- 0). As we shall see in Chapter IV, this does not through (x0,yO,Yo,O), both hypotheses (H10) and (H11) of §4 are violated. In this case,iJ:is necessary to find a way to "factor out" a branch so that the methods of §4 and §5 apply. One is then faced with determining how the branches intersect. we shall demonstrate how this may be done in an application to chemical reaction equations discussed in Chapter IV. 71 Q6. Solutions near (XO’YO’YO’O) = (0,0,0,0) In this section, we shall informally discuss some ways of determining the behavior of solutions of (4.1) near (0,0,0,0). At (0,0,0,0), the matrix and does not have rank two. Thus the hypotheses (H9)-—(HlO) of §4 do not apply. The system F n n-l n (6.1)(a) o = F(x,y,y,§) = Tx+anx +an-1x Y+"‘ +‘J‘oY + §Nl (x: Y: V) g) n-l n n (6.1)(b) O G(x,y,y,§) = y-tbnx 'kbn-lx y-+°°'-+boy k + §N2(X:Y:Y:§) is similar to the l-dimensional case of Chapter II as y passes through 0 at g = 0, except for the extra parameter 1;. Example 6.1. Consider the system 2 0 ny-x (6.2) 0 = Tx-Tnxz-Tyz. After the previous scaling, we have 0 y-—x (6.3) 0 = YX+T1X2 +y2 3 T YT)- 72 Solutions near (x,y,y,n) = (0,0,0,0) correspond to solu- tions which bifurcate from the n-axis near (0,0). If (x,Y) is a nontrivial solution, elimination of y shows that x must satisfy 0 = y+fix+x3 The discriminant of this cubic is given by —4n3-27Y2. Thus along the curve 4n3 = ~27y2 in the (y,n)-plane (or 4fi5 = —27T2 in the (T,n)-plane) the number of non- trivial solutions changes by two. A ”‘4 Figure 10. Behavior of nontrivial solutions of system (6.2). Figure 10 shows the projection of the "solution space" unto the (T,n)—p1ane. This example shows the advantage of con— sidering a full neighborhood of (T,n) = (0,0) in our 73 studies. It does not suffice to fix n and vary T to study solutions which bifurcate from the n-axis. This example shows that we may eXpect to find a cusp on curves tangent to the axes in the (T,n)—plane. These curves and cusps play a role similar to the fold curves discussed earlier. Since FY = l at (0,0,0,0), we may use the implicit function theorem to solve for a unique y(x,y,g) such that F(x,y(x,y,§),y,§) = 0 and y(0,0,0) = 0. Since (0,0,y,§) is a solution for all (y,g) sufficiently close to (0,0), y(O,Y,§) = O and y(x,y,g) = xH(x,y,g) where H is ana- lytic in x, y and g. Substitution of y(x,y,g) in F1, and division by x gives an equation of the form (6'4) 0 = F(X:Y:§> = Y+[f1(Y:§)]X+[fZ(Y,§)]x2+~~ and we have successfully "divided out" the solution (x,y) = (0,0). At this point we may take the following approach: solve (6.4) for y(x,g) and analyze the singularity of the map T: (x,§) 4 (y(x,§),g) at (x,§) = (0,0). We shall sketch how this may be done for the following special case (6.5) (a) O yx-Talx2-+b1xy-+cly2-+§Nl(x,y,y,§) (6.5) (b) o 2 2 y-Tazx -+b2xy-+c2y -+gN2(x,y,y,g) 'We now find that 2 3 y(x,y,§) = -a2x -+b2a2x -+xH(x,y,g). 74 substitution of y(x,y,§) in (6.5)(a) and division by x yields, _ 2 3 where fl(0,0) = a1, f2(O,O) = -bla2, and f3(0,0) b1b2a2-c1a2. If a1 #’O, we may solve for a unique X(y,§). Since ala2 = 0 would make the bilinear form F 2 2 alx -+blxy-+cly M((X:Y): (X:Y)) =< 2 2 Lazx -+b2xy-+c2y degenerate, we now assume that a1 = 0, a2 #’0. As before, we may solve for y(x,§). ‘we find that %%-= a1 = O and 2 h—% = —2b1a2, at (x,g) = (0,0). If b1 #'0, there is a BX 3 "fold" [ 3] tangent to the axis. If bl = 0, i—% = -6c1a2. 2 Y" If c1 # 0, and in addition,-§§§% #’O at (x,§) = (0,0), there is a cusp [ 3] at (0,0). The situation is similar to Example 6.2. 'We now outline an alternative approach which is more analytic in nature and does not simply entail applying class- ifications from singularity theory. This is useful for problems which do not fit the standard classifications. The main idea is to find curves (x(g),y(§),y(g),§) such that x(0) = y(0) = y(0) = O and the Jacobian of system 75 (6.1) has zero determinant along these curves. These curves are determined by the system r0 =F1(X:Y(X)Y:§)1Y)g) = Yx+fl(Y:§)x2+f2(Y)§)3+"° (6.6% K0 = §F1(X,Y(X, Y: ELY: g) = Y +2fl(Y: §)X + 3f2 (Y: §)X2 +’ ' ' since a}? as a = ._1 ___l. ax = axF1(x’Y(x’Y’§)’§) ax + By ax aFl 5F aFZ AFZ =__ 1(-_ ax ay. ax ay =0 at (x,y(x,y,§),y,g) if and only if 1- ‘1 asl aFl ax ay det = 0. 532 .532 t. ax ay .— One such curve is x(g) = 0 = y(g). This corresponds to the g-axis and the trivial solution (x,y) = (0,0). We are interested in curves where x(g) #’O for g #'0. Let xG(x,y, g) = Fl(x,y(x,y,g),y, g). G(x,y,§) = 0 implies 523141 (X,Y(X,y,§),v.§) = X%G(X.Y,§). Thus nontrivial solutions of (6.6) are found by studying (6.7)(a) O = G(x,y,g) = y+f1(y,§)X+--- (6.7)(b) o = 35,23 (X,y,§) = f1(Y:§)+2f2(Y:§)X+'°' 76 Since %§-#'0 at (0,0,0), we may use the implicit function theorem to solve for y(x,§) with y(0,0) = 0. y(x,§) is analytic in (x,g) for (x,§) in a neighbor- hood of (0,0). Substitution of y(x,§) in (6.7)(b) gives a power series C(x,§) which converges in some neighbor- hood of (0,0). The curves x(g) may be determined by Newton's polygonal method. CHAPTER IV APPLICATION TO A SYSTEM OF CHEMCIAL REACTION EQUATIONS 91. Introduction In this chapter we shall apply the method of the pre- ceeding chapter to a system of equations that arises in the study of chemical reactions [ 2]: 2 (1.1) 33-:A-(B+1)x+x2y+D-a—’5, O0 3r with boundary conditions X(O,t) = y(l,t) =A, y(o,t) = y(l,t) B/A. A and D are constants, and D = Dx’ vD = DY’ where Dx and Dy are diffusion coefficients for x and y respectively. B and v are taken to be the parameters. For all values of B and v, x0 = A, y0 = B/A is an equilibrium state. Substitution of x = An+u(r,t), y = B/An+v(r,t) into (1.1) gives 77 78 2 .52: [3-1]u+A2v+D3—1—1+§u2+2Auv+u2v at arz A (1. 2) 2 A! = -Bu -A2v +\)D L2: —-B-u2 - 2Auv - u2v A ar with boundary condition u(0,t) = u(l,t) = v(0,t) = v(l,t) = 0. The associated linear system is f' 2 53 = [B-l]u+A2v+D L-u- at 612 (1.3) < 2 31:- _ 2 5.1 Lat Bu AV+VD 2 hr with the same boundary conditions. Since we are interested in steady state solutions, we will consider u and v as functions of r, and study the stationary problem F 2 O = [B- l]u+A2v+D 5L; + guz + 2Auv+u2v dr (1.4) 2 0 = -Bu-A2v+vD 534% -% 2- 2Auv-u2v k ar with boundary conditions (1.5) u(O) = v(O) = u(1)= v(l) = 0 Let Y be the Banach space of twice continuously dif- ferentiable functions from [0,1] to IR2 which satisfy (1.5) and 79 (1.6) U”(0) = V” (0) = l1”(l) = V”(l) = 0. Let 1(u:v>1y= 111111+11v11+11u'11+11v'11+11u”11+11v”11 where “-H denotes the sup norm. Note that if (u,v) is a solution of (l.4)-(l.5), then (1.6) is automatically satisfied. Let 2 7 ' T (B-l)+DL A2 u A(V,B)[u,v] = dr 2 -B -A +vD-g—‘2- V - dr .1 - .1 and ~ 1 l 2 BA- u + 2Auv + uzv N(B,u,v) = L-BA_lu2-2Auv - uzv‘ If (u,v) is in Y, then A(vgB)[u,v] and N(B,u,v) satisfy (1.5) for all v and B. Let X be the Banach space of continuous functions from [0,1] to JR2 which satisfy (1.5). Let [(u,v) )x = ”u“ + “V“ Then A(v,B): Y 4 X. is a continuous linear operator for each (v,B) in 1R2, and N: IRx Y 4 X is continuous. The stationary problem (l.4)-—(l.5) is equivalent to (LH o=Awmnmw+mhmw. 80 If we replace R2 by C2 in the above, we see that .A(v,B) and N have continuous complex extensions. A(v,B) depends on v and B analytically and N depends on B, u, and v analytically. §2. Eigenvalues, Eigenvectors and Projections The spectrum of the linear operator A(v,B) consists of the eigenvalues + (2.1) a; =-§1-[B-l-A2-n21r2D(l-1-v) : [[B -1 +112 +n21T2D(\) - 1)]2 -4A2B]1/2}. The corresponding eigenfunctions are sinrmm: + (202) Q; = + . M; Sinlnn: + where M; satisfies + (2.3) 0;'- (B-l)-+n2w2D - A2 = 0. 5l+ + For real 05, the projection unto the linear subspace spanned + by 6; is given by 1 I u .i _ 2 . —- . .i (2.4) Pn[v:)§n - ii J‘ us1nn7rr + Nn vs1nn1rrdr 6n 1-+M N 0 n n i where Nh satisfies + 2 2 .i (2.5) o;- (B-1)+n7rD+BNn=O. 81 sin mrr (2.65) UI+ .i N sin n'rrr n is the solution of the adjoint equation YT ’YB-l)+D-JL -B 7 VA 1 2 1 (2 7) '1 ar . o' = n 2 )1 A2 -A2 + vD 3— ‘11 2 2 2 L _, L_ ar _, L. _a with boundary condition Y1(O) = Y2(0) = 1111(1) = Y2(l) = 0. + + We can see that P;[ 6;] = 1. We shall now determine a curve in the real (V,B)-plane such that all the eigenvalues of A(B,v) have negative real part when (v,B) lies below this curve [ 2 ]. This curve will be called the grave 9_f_ neutral stabilifl. We can see that Rec; 2 Red; for all n. If 0 + . n is complex, then the curve Red; = O is given by the straight line (2.8) 2 B = 1+A +n2'11'2D(1+V). If 0; is real, Rec; = O is given by the hyperbola 1%. (29) B-l+n22D+Ail+——l—D o — 1T \){ 2 2 30 n 1r . + A typical curve Reo '- n — O is given in Figure 11. 82 + Figure 11. The curve Reon = 0. The curve we seek is found by joining portions of the curves ‘Reo; = 0 which lie lowest for each (v,B). Red: = 0 lies lowest for sufficiently small v and for sufficiently large \n For values in between, a finite number of the hyperbolas lie lowest. A typical curve is "scalloped" and is given by Figure 12. V Eimnelz. A typical curve of neutral stability. 83 'mhe intersection of Hn and Hn+1 is given by A2 n2 (n+1) 2 (T212) 2 B = l-+n2w2D-+n2(n-+l)2(w2D)2[1 + 212 } n W D 2 2 = l-Tn W 2 D-+(n-+l)zw D-l-n2(n-+1)2(1T2D)2 2 = (l-tn WZD)(l-+(n-+l)2w2D) The number of such intersections which lie on the curve in Figure 12 depends on the values of A and D. we will study bifurcation in the neighborhood of (Vc’Bc) for n 2.1. Proposition 2.1. If 0; # a; for n ¢'m is real, then it has multiplicity 1. That is, if [A(B,v)-o;I]k§ = 0, + then s = K§n' Proof. We proceed by mathematical induction. It is true for k -1. Suppose [A(B,v) -o;I]k§ =0. Let 3 = + k-l .. + + + [A(B,v)-onI] Q. Then 6 = cgn and Pn[cqn]§n = Pn[[A(B,v)-U;I]k-l§]§;. Since Pn commutes with A, c4; = [A(B,\)) -O;I]k_1[Pn[¢]§;] = 0. Therefore c 0 and the induction hypothesis implies + 6 - KQno Thus we have exactly the situation described in Chapter III. 84 §3. Calcuation of the Bifurcation Equation Let (u,v) in X be decomposed as follows. u sin nTrr sin (n+l)7rr W1 (3. l) = x + y + v Mn 811') mm Mn+l Sln (n+l)7rr sz = xgn + Y§n+l + w where u u w1 x = Pn , y = Pn+l , and w = 18 in Qx. v v w2 M N M N c rres ond to 0+ = O - 0+ and n’ n’ n+1’ n+1 o p n - n+1 are given by (3.2) Mn = -A‘2[Bc+1-n21rzn] (3.3) Mn+1 = -A'2[Bc+1 - (n+l)21r2D] (3.4) Nn = B;1[Bc-l-n21r2D] (3.5) NM1 = B;1[Bc-1- (n-l)21r2D]. The subspace Qx is given by wl w {(wlwz) e XIPn ll '1! II 0 9...) n+1 w2 w2 <2Y = Ynox. 85 The auxiliary equation is given by (3.6) o = [I-Pn-Pn+1][Aow+(B-BC)A1(u,v)+(v-vc)A2(u,v)] + N(B,u,v) where A0 = A(VC,BC), 1 O O 0 A1 = ’ A2 = dz -1 O 0 ‘——5 dr A-lu2 and N(B,u,v) = N(B ,u,v)-+(B-B ) . c c -A-1u2 Proposition 3.1. There exists a unique real analytic function w(x,y,v,B) such that XQD'FY§n+1'*W(X,Y,v,B) is a solution of (3.6) for all (x,y,v,B) in a neighborhood of (0,0,0,0) and w(0,0,0,0) = 0. Moreover, w(x,y,V,B) can be expressed by a convergent power series in this neigh- borhood. [This power series contains terms of second order or higher in x,y,v,B and all terms contain at least an x or y.] Proof. we need only show that the restriction (3.7) A is a linear homeomorphism, and then apply the implicit func- tion theorem. The rest follows from the remarks of I§3 and 86 equating coefficients when the power series for w(x,y,v,B) is substituted into (3.6). Since there are no eigenvalues of H0 in Qy, the map (3.7) is one-to-one. It is onto by a "Fredholm alternative" type theorem for boundary value problems [‘9]. Hence by the open mapping theorem it is a linear homeomorphiamdj The bifurcation equations are given by (3.8) (a) 0 Pn[(B-BC)A1(U,V) + (v-vC)A2(U,V) +N(B,u,V)] (3.8) (b) o = pn+1[ (B-Bc)Al(u,v) + (v 'Vc)A2(“’V) +N(B,u,v)]. Substitution of w(x,y,v,B) into (3.8) gives a system of the form 0 = F(X,Y,\),B) 0 = G(X,Y,V,B) where F and G are power series that converge in some neighborhood of (0,0,0,0). From this point on, we shall assume that n is odd. F and G are actually the coefficients of § = ' and Q +1 = O O W 8111 mrr 1 8111 (n + l)1rr ' - n Mn 81!) mrrj J Mn sin (n + l)1rr respectively. 'We Shall demonstrate how symmetry properties can be used to obtain information about the bifurcation equations. 87 Recall the original system for the stationary problem 0 = A(B,v)[u,v] +N(B,u,v) . Let 9 be the transformation given by e(u(r),v(r)) = (u(l-—r),v(l-—r)). It is easy to see that g is an involution and that it commutes with the operators A(B,v) and N(B,u,v) for all v and B. ‘We have 9 (Mn +Y§n+1 +w) XQn - ”n+1 + 9w and QW(XJY:V:B) = W(X, 'Y:V:B)° Thus F and G must satisfy F(X,-Y,\),B) F(X,Y,\),B) -G(X,-Y,V,B) G(XJY:V:B)’ for all x,y,v,B in some ball about (0,0,0,0), and we have the following. Proposition 3.2. If n is odd, the bifurcation equa- tions (3.2) can be reduced to the form (3.9)(a) O = F(x,y,v,B) = Tx-+ax2-+by2-+(higher order terms) (3.9) (b) O G(x,y,v,B) ny-+cxy-+(higher order terms) 88 where n = n(v,B), T = T(v,B). Here F(x,y,v,B) contains only even powers of y and G(x,y,v,B) contains only odd powers of y. As in the previous chapter, we may show that the first terms of (3.9)(a) are + + on aon '?fi§(vc’Bc)(B"Bc)x + ifir(vc’Bc)(V"vc)x'+°°' . Those of (3.9)(b) are + 80 ac —a‘;3-‘fl = cxy L is nondegenerate, and we may use the change of scale given in III §3. Recalling the consequences of symmetry given in proposition 3.2., the bifurcation problem is equivalent to studying the two systems (3.ll)(a) O = x-+ax2-+by2-+TF1(x,y2,y,T) (3.11) (b) o = yy+cxy+TyF2(x,y2,Y,'r) and (3.10)(a) 0 = yx-+ax2-+by2-+nGl(x,y2,y,T) (3.10) (b) 0 = y +cxy +71sz (x, yzm, 'r) where all terms in F1, F 61’ 62 are second order or 2) higher terms in x and y. 92 If (x,y,y,T) is a solution of (3.11), then (Tx,Ty,T,yT) is a solution of (3.10). If (x,y,y,n) is a solution of (3.12), then (nx:nY:Yn:n) is a solu- tion of (3.10). Equation (3.11) will be used to study solutions of (3.10) corresponding to (T,n) in a set of the form S1 = {(T:fi)| |T‘ 2.p1|n‘} where pl > O. Equa- tion (3.12) will be used for sectors 82 = {(T,n)\ 'n‘IZ PZIT‘}- S1 and 82 are given in Figure 13, Figure 13. The sectors S1 and 82. (Of course p1 and p2 have been chosen so that the sectors overlap). 94. Analysis of the Bifurcation Equations Recall that n is odd and that we assume A and D are chosen so that a, b and c are nonzero. 93 When T = O, (3.11) becomes 0 = x-i-axz-i-by2 (4.1) O yy + cxy which has the nontrivial solutions 1 _ (4.2)(3) X=--a-, y—O and y(1 -%Y) (4.2)(b) x=-3cfi, y=: cb . These solutions coincide when y = c/a. The matrix of (H10) at a solution (x0,yo,yo,0) is given by b r 7 l + 2axO 2byO 0 Fl (x0, yo, YO’ O) L Cyo Y+cxo yo F2(xo’yo’Yo’0U . _ 2 The determinant J(xo,yo,yo,0) — (l-+2ax0)(y+cxo)--2bcyo 1 3.903%: 0) and (030,030) is zero only for the solution (- at T = 0. Thus if (xo,yo,yo,0) is any other real solu- tion, there are unique x(y,T), Y(Y,T) such that x(yo,0) x0, y(yo,0) = y0 and (X(y,T),y(y,T),y,T) is a solution of (3.11) in some neighborhood of (yo,0). It remains to determine the behavior of real solutions of (3.11) near (_ 1 C 530,530) and (0,0,0,0). At the solution (- %, O,n§,,0), the matrix given in (H11) becomes 1 l O O O O J J —l J L X Y T J and therefore does not satisfy the hypotheses (H10)-(Hll) of Chapter III. However, we shall still be able to decribe l the behavior near (- 3, O , g, 0) . Recall that the bifur- cation equations have the form (4.3) (a) 0 x +ax2 +b)/2 +TF1(X:Y2: Y, T) (4.3)(b) o y(y+cx+TF2(x.yz.y.T)) after scaling. ‘We may apply the implicit function theorem to (4.3)(a) to obtain: Proposition 4.1. There is a unique solution branch of (4.3) of the form (x(y,T),O,y,T) in a neighborhood of (— i, O , g, 0), with x(g, O) = - i. This branch is deter- mined by (4.3)(a) and y = 0. We shall refer to this branch as branch I. We shall see that another branch is determined by (4.4) (a) O 2 2 2 x+ax +by +TFl(x,y ,y,'r) (4.4) (b) o Y'+CX'*TF2(X,Y2:Y,T) 95 Let 3(x,y,y,T) denote the Jacobian of (4.4) at (x,y,y,T). The matrix corresponding to (4.6) in Chapter III is given by r- '1 -1 O O c 0 l O 2bc O L e and has nonzero determinant. By Theorem III there is a unique fold curve §(T), with corresponding solutions (anion) such that $40) = -§-, Em) = - i, 37(0) = o, o = 3(i(7),§(T),§(T),T), and(4.4) determines a solution 1 c branch near (- 3, O, 3, O). The direction of the fold is determined by the number of real solutions of (4.4) for T = O and y near Eu These solutions are given by (4.2)(b). we have the follow- ing. Proposition 4.2. The system (4.4) determines a solu- tion branch of (4.3) which satisfies (a) If be > 0, there are two real solutions of (4.4) for Y < Y(T) and none for Y > Y(T). (b) If bc < 0, there are two real solutions of (4.4) for y > Y(T) and none for y < Y(T). Note that this result is local and only holds in a 9,0). neighborhood of (x,y,y,T) = (--%,,O, a 96 We shall refer to these solutions as branch II. Branch II is determined at T = 0 by (4.2)(b) and in this sense is distinct from branch I which is determined by (4.2)(a). l c At (--;, O, 3’ O), brandh I and branch II coincide. Proposition 4.3. The only intersection of branches I and II is the set of solutions (2(T),§(T),§(T),T). Proof. For any (x,y,y,0), we have 1+2ax+T§ 2by+TW J(X: Y: YJO) = det 5F2 3F2 = O + —— _— .. C T ax T ay 4 when y = 0, because —a-y—(X,O, Y,T) = O = $090, Y, 1') = 0. Since (4.4) determines x(T) and y(T) uniquely when y = O, the solution corresponding to the unique fold curve §(T) has the form (imp) with x(0) = - %. Hence (x(T),y(T),§(T),T) is part of branch I. Secondly, the two branches can intersect only when the Jacobian J(x,y,y,T) of the system (4.3) is zero and y = O. 'we have _ BF]. .1 1+2ax+T —— 0 5X J(x,0,y,T) = det 2 _ O y+cx+TF2(x,y ,Yfl-L. 97 we see that J(x,0,y,T) = O at a solution (x,0,y,T) of (4.3)(fbr 1'sufficient1y small), if and only if (x,0,y,T) is also a solution of (4.4). There is a unique solution of (4.4) of this form and it is the one corresponding to the fold curve. The situation just studied is completely different from that considered in Chapter III. we may view branch II as bifurcating from the solution (i(t),0) of branch I at (7%), T). we now determine the behavior of real solutions of (4.3) near (0,0,0,0). Proposition 4.4.(a) If cb > O, and T and y are sufficiently small, there are two nontrivial real solutions (x(y,T),y(y,T)) of (4.3) for y'> 0 such that x and y tend to zero with y. There are none near (0,0,0,0) for y(O. (b) If cb < 0, there are two nontrivial real solutions for y < O, and none for y > 0. At T = 0, these solutions are given by (4.2)(b). Proof. we may apply the implicit function theorem to (4.3)(a) to obtain x(y,y,T) such that (x(y,y,T),y,y,T) is a solution of (4.3)(b) in a neighborhood of (0,0,0) and x(0,0,0) = O. substitution into (4.3)(b) gives an equation of the form 98 3 3 o = YY-C'bY +y g(y,Y,T) where g(0,0,0) = 0 [Terms in g(y,Y,T) which do not con- tain T or y as a factor will be at least first order in y]. Division by y takes care of the trivial solution that exists for all y. The results follow from a consider- ation of 2 2 (4.5) 0 = Y—cby +y gum/n) for T and y sufficiently small.D In fact, the curve Y(T) = O is a fold curve for the system consisting of (4.3)(a) and (4.5), and the solution correSponding to this fold curve is (x(T),y(T)) = (0,0). The line Y(T) = 0 corresponds to the T-axis in the T’fi plane. Thus branch II bifurcates from the trivial solution at the T—axis, and disappears at the curve of solutions (T§(T),O) of branch I when y reaches ?(T); i.e., (T,“) reaches (T,?(T)T) in the T-n plane. In the (T,n)- coordinates, branch II solutions have the fbrm x ll —TY/c+ooo =—n/c+ooo /y 11 -a/cY1 +cb/nU -a/cn) y=irr Cb +ooo=_ +... All that remains is to determine the behavior of solu- tions for Y near zero in (3.12). [This corresponds to Y near "a“ in (3.11); or the n-axis in the T-n plane]. When T = 0, (3.12) becomes 99 O = Yx-i-axz-i-by2 (4.6) O y-Tcxy Nontrivial solutions of (4.6) are given by (4.7) x=-§. y=o and (X-C/afi (4.7)(b) X=-%, y=i/ be The solutions (4.7)(b) are real for y near 0 if and only if bc and c/h have opposite signs. The solu- tion (4.7)(a) approaches the origin as Y tends to zero. Proposition 4.5. There is exactly one nontrivial solution (x(Y,n),y(y,n)) of (4.6) in a neighborhood of the origin, such that x(Y,n) and y(y,n) tend to zero with y. This solution is given by (4.7)(a) when n = 0. Proof. As in proposition (4.4), y(x,Y,n) is deter- mined by (3.12)(a) and the problem reduces to 2 2 (4.8) 0 = Yx+ax +x h(x,y,n) where h(0,0,0) = 0. Since a #’O, the result follows from the implicit function theorem.U Thus branch I bifurcates from the trivial solution at the n-axis. In (T,n)-coordinates branch I solutions have 100 the form x = - Yn/é-t--o = -T/h'+°" : Y = 0- We now choose the sectors S1 and S2 mentioned in §3. Let S1 = [(T,n)| |T\ 2.91ln\} ‘where 0 < p1 < [a/c|. p2 is chosen so that it overlaps 81, but does not contain the line n = (c/a)T. we summarize our results by the following bifurcation diagrams in the (T,n)-plane (Figure 14). The numerals de- note the number of nontrivial solutions of (3.10) which correspond to (T:n) in each sector. The curve §(T) given by proposition (4.2) has become the curve F(T) = (T,§(T)T) which passes through the origin with slope c/a. lOl h n F I \f\\ \\\\ . - 1e ‘ \\\ 1" f 17 F A Case 1. bc > O, c/a > 0 Case 2. be > O, c/a < O n “7, F ”1 Case 3. be < O, c/a > 0 Case 4. bc < 0, c/a < 0 Figure 14. Bifurcation diagrams for the system (3.10) in the (T,n)-plane. Figure 15 is an attempt at a more geometrical repre- sentation of the situation. We are only considering Case 4 in Figure 15. 102 B Intuitive representation of bifurcation and c/a < O. 103 Figure 15 is meant more as an aid to the intuition rather than an actual representation of the solutions (x,y,T,n). One must interpret the drawing in the following spirit. Let T2-+n2 = c be a sufficiently small circle in the T‘fi plane. As we pass through point A, solutions of branch I (broken line) bifurcate from the trivial solution at the n-axis. As we move around the circle counterclockwise the corresponding solution of branch I passes through branch II at B, proceeds to bypass the trivial solution at the T-axis, and finally passes through the trivial solution when it reaches the n-axis again at C; etc. Solutions of branch II (double valued) bifurcate from the T-axis and continue until they disappear into branch I at the solution (T§(T),O) corresponding to r. [1] [2] [3] [4] [5] [5] [7] [8] [9] [10] [11] BIBLIOGRAPHY Andronov, A.A., E.A~ Leontovich, I.I. Gordon, and A. G. Maier, Theory Lf Bifurcations of Dynamic Systems Ln a Plane, John Wiley & Sons, Inc., New Ybrk, 1973. Boa, J.A., and D.S. Cohen, Bifurcation of Localized Disturbances in a Model Biochemical Reaction, SIAM J. Appl. Math., 30, no. 6 (1976), 123-135. Br5cker, T.H., and L. Lander, Differentiable Germs and Catastrophes, Cambridge university Press, Great Britain, 1975. Chow, S.N., J.K. Hale, and J. Mallet-Paret, Appli- cations of Generic Bifurcation I, Arch. Rational .Mech. Anal. 59 (1975), 159-188. [Applications of Generic Bifurcation II to appear]. Dieudonné, J., Foundations 9§_Modern Analysis, Academic Press, New Ybrk, 1960. Gavalas, G.R., Nonlinear Differential gguations Lf Chemically Reacting Systems, Springer, New York, 1968. Hale, J. K., _pplications Lf Alternative Problems, Lecture notes 71- 1, Division of Applied Mathe- matics, Center for Dynamical Systems, Brown University, Providence, R.I. Hale, J.K., Ordinary Qifferential Equations, in "Pure and Applied Mathematics", Vbl. XXI, Wiley- Interscience, New YOrk, 1969. Hartman, P., Ordinary Differential Equations, John ‘Wiley & Sons, Inc., New York, 1964. Henry, D., Geometric theory of semilinear parabolic equations, unpublished lecture notes. Kato, T., Perturbation Theory for Linear Operations, Springer-Verlag, New YOrk, 1966. 104 [12] [13] [14] [15] [161 [17] [18] [19] 105 Keller, J.B., and S. Antman, Editors, Bifurcation Theory and Nonlinear Eigenvalue Problems, Courant Institute of Mathematical Sciences, New YOrk University, New Ybrk, 1967. Kirchggssner, K., and Sorger, P., Stability analysis of branching solutions of the Navier-Stokes equa- tions, Proc. TWelfth Internat. Congress Appl. Mech., Stanford, 1968, 257-268. Kirchgassner, K., and H. Kielhcfer, Stability and bifurcation in fluid dynamics, Rocky Mtn. J. Math., 3, no. 2 (1973), 275-318. Nachbin, L., pr01092 22.Spaces g; Hblomorphic Mappings, Springer-Verlag, New Ybrk, 1969. Rabinowitz, P.H., Existence and nonuniqueness of rectangular solutions of the Bénard problem, Arch. Rational Mech. Anal., 29 (1968), 32-57. Sattinger, D.H., Stability of bifurcating solutions by Levay-Schauder degree, Arch. Rational Mech. Anal., 43 (1971), 154-166. Yudovich, V.I., On the origin of convection, PMM (Jour. Appl. Math. and Mech.) 30, no. 6, (1966), 1193-1199. Yudovich, V.I., Free convection and bifurcation, PMM 31, no. 2, (1967). HICHIGRN STGT u [[1 [g 3 2 31 E UNIV. LIBRnRIES [IIIIIWIWIIHIIIHHINIIW\l 05309474