On the numerical solution of certain nonlinear systems arising in semilinear parabolic PDEs? M. De Leo1 , E. Dratman2 , G. Matera3 1

Departamento de Matem´ atica, FCEyN, UBA, Ciudad Universitaria, Pabell´ on I (1428) Buenos Aires, Argentina. [email protected] 2 Departamento de Computaci´ on, FCEyN, UBA, Ciudad Universitaria, Pabell´ on I (1428) Buenos Aires, Argentina. [email protected] 3 Instituto de Desarrollo Humano, Universidad Nacional de General Sarmiento, J.M. Guti´errez 1150 (1613) Los Polvorines, Buenos Aires, Argentina. Member of the CONICET, Argentina. [email protected]

Abstract. We consider a family of polynomial systems which arises in the analysis of the stationary solutions of a standard discretization of certain semilinear second order parabolic partial differential equations. We prove that this family is well–conditioned and exhibit a polynomial– time homotopy continuation method solving any member of this family, significantly improving the exponential behaviour of all known general algorithms solving a generic instance of this family.

1

Introduction.

Several scientific and technical problems require the solution of multivariate polynomial systems over the real or complex numbers (see e.g. [20], [17]). Numerical methods computing all solutions of a given 0–dimensional square polynomial system f1 = 0, . . . , fn = 0 usually rely on deformation techniques, based on a perturbation of the original system and a subsequent path–following method (see e.g. [16], [1], [13], [14], [3], [24]). Such kind of methods are usually known as homotopy continuation methods, and have a complexity of LnO(1) Dµ2 floatingpoint operations, where – n is the number of variables of the input system, – L is the number of arithmetic operations required to evaluated the input polynomials f1 , . . . , fn , – D is the number of paths to be followed, – µ is highest condition number arising from the application of the Implicit Function Theorem to the points of the paths produced by the homotopy. Let us observe that the parameters L, n are determined by the input equations, while D is usually estimated by a suitable B´ezout number associated to the ?

Research was partially supported by the following Argentinian and German grants: UBACyT X198, PIP CONICET 2461, BMBF–SETCIP AL/PA/01–EIII/02, UNGS 30/3005 and beca de posgrado interno CONICET.

structure of the problem (see e.g. [18], [11], [10]). Taking into account that L and n are generally much smaller than D and µ, we conclude that the complexity of an homotopy algorithm is essentially determined by the parameters µ and D. Following [3], we shall call the input system well–conditioned if the parameters O(1) µ, D are . For a generic input system one has Qn of kind max1≤i≤n {deg(fi )} O(1) D = i=1 deg(fi ) and µ = max {deg(f . Furthermore, worst-case 1≤i≤n i )} Qn O(1) estimates for µ are also of kind i=1 deg(fi ) , and it is not possible to decide a priori whether the condition number µ of a given system has such an exponential behaviour. As a consequence, standard homotopy continuation methods require Qn an exponential complexity L(n i=1 deg(fi ))O(1) , which is unsuitable for real-life purposes. In this article we develop an efficient method to compute the positive solutions of a particular family of polynomial systems, which arises from a discretization of certain second order parabolic semilinear equations. General homotopy continuation methods applied to this family have the exponential complexity behaviour mentioned above. Let be given univariate rational polynomials f, g, h and consider the following initial boundary value problem:  ut = f (u)xx − g(u) in (0, 1) × [0, T ),    f (u)x (1, t) = h(u(1, t)) in [0, T ), f (u) (0, t) = 0 in [0, T ),  x   u(x, 0) = u0 (x) ≥ 0 in [0, 1]. This kind of problems models many physical, biological and engineering phenomena, such as heat conduction, gas filtration and liquids in porous media, growth and migration of populations, etc. (cf. [9], [21]). In particular, the long–time behaviour of its solutions has been intensively analyzed (see e.g. [5], [12], [23]). The usual numerical approach to this problem consists of considering a second order finite difference discretization in the variable x, with a uniform mesh, keeping the variable t continuous (see e.g. [2]). This semi–discretization in space leads to the following initial value problem:  ¡ ¢ u01 = 2(n − 1)¡2 f (u2 ) − f (u1 ) − g(u1 ),   ¢  u0k = (n − 1)2 f¡ (uk+1 ) − 2f (uk ) ¢+ f (uk−1 ) − g(uk ), (2 ≤ k ≤ n−1) (1) u0n = 2(n − 1)2 f (un−1 ) − f (un ) − g(un ) + 2(n − 1)h(un ),    uk (0) = u0 (xk ), (1 ≤ k ≤ n) where x1 , . . . , xn define a uniform partition of the interval [0,1]. In order to describe the dynamic behaviour of the solutions of (1) it is usually necessary to analyze the behaviour of the corresponding stationary solutions (see e.g. [4], [8]), i.e. the positive solutions of the following polynomial system:  ¡ ¢  0 = 2(n − 1)¡2 f (u2 ) − f (u1 ) − g(u1 ), ¢ 0 = (n − 1)2 f¡ (uk+1 ) − 2f (uk ) ¢+ f (uk−1 ) − g(uk ), (2 ≤ k ≤ n − 1) (2)  0 = 2(n − 1)2 f (un−1 ) − f (un ) − g(un ) + 2(n − 1)h(un ). A typical case study is that of the heat equation, i.e. f (X) := X, with nonlinear reaction and absorption terms (see e.g. [4], [5]). In this article we shall consider

the case where h is a constant, and g(X) is a polynomial of Q[X] with g(0) = g 0 (0) = 0, g 0 (x) > 0 and g 00 (x) ≥ 0 for any x > 0, i.e. the initial boundary value problem  ut = uxx − g(u) in (0, 1) × [0, T ),    ux (1, t) = α in [0, T ), (3) in [0, T ),  ux (0, t) = 0   u(x, 0) = u0 (x) ≥ 0 in [0, 1], and the corresponding set of stationary solutions of its semi–discretization in space, i.e. the positive solutions of the following system:   0 = 2(n − 1)2 (u2 − u1 ) − g(u1 ), 0 = (n − 1)2 (uk+1 − 2uk + uk−1 ) − g(uk ),  0 = 2(n − 1)2 (un−1 − un ) − g(un ) + 2(n − 1)α.

(2 ≤ k ≤ n − 1)

(4)

Such kind of systems have been considered in [7], [6]. In these articles it is shown that (4) has only one positive real solution, and a numeric algorithm solving particular instances of (4) with O(Ln9 ) operations is proposed. Here, we significantly improve and generalize this approach, by exhibiting an algorithm which solves (4) with roughly O(Ln3 ) operations. Let us remark that standard numeric algorithms solving (4) require an exponential number of floating-point operations L deg(g)O(n) . In Section 2 we prove that the solutions of the semidiscrete version of (3) converge to the corresponding solutions of (3) in any interval where the latter are defined, showing thus the consistence of our semi–discretization. We further show that any solution of the semidiscrete version of (3) which is globally bounded converges to a stationary solution of (3). Then, in Section 3 we exhibit an homotopy deforming (4) into a triangular system, with only one positive solution, which is easy to solve. We prove that the homotopy path which deforms the positive solution of (4) into the positive solution of this triangular system is well-conditioned. This allows us to exhibit an homotopy continuation algorithm which computes an ε–approximation of the only positive solution u∗ of (4) with O((L+n)(n2 +M )) floating point operations, with M = O(log | log(εn)|). The starting point of this algorithm is the positive solution of the triangular system produced by the homotopy, and hence it does not depend on random or generic choices.

2

The Initial Boundary Value Problem under Consideration.

As mentioned in the introduction, we shall consider the initial boundary value problem (3) for u0 (x) satisfying the “compatibility condition” u00 (1) = α, u00 (0) = 0, and g being a polynomial of Q[T ] with g(0) = g 0 (0) = 0 and g 0 (x) > 0 for any x > 0 (for example g(X) = X d ). In order to solve (3), we consider the following

(semi)discrete version of (3):  0 ¡ ¢ ¡ ¢ u1 (t) = 2h−2¡ u2 (t) − u1 (t) − g u1 (t) ,¢   ¡ ¢  0 uk (t) = h−2 u + uk−1 (t) − g¡ uk (t)¢ , (2 ≤ k ≤ n − 1) k+1 (t) − 2uk (t) ¢ ¡ u0 (t) = 2h−2 un−1 (t) − un (t) + 2h−1 α − g un (t) ,    n uk (0) = u0 (xk ), (1 ≤ k ≤ n)

(5)

where x1 , . . . , xn define a uniform partition of [0, 1] and h := (n − 1)−1 . We are going to show that the solutions of (5) converge to the corresponding solutions of (3), and we shall discuss the role of the stationary solutions of (5) in the description of the asymptotic behaviour of the solutions of (5). We start with the convergence result: Theorem 1. Let 0 < τ ≤ T be a value for which there exist a unique positive solution u(x, t) ∈¢ C 4,1 ([0, 1] × [0, τ ]) of (3) and a numeric approximation U (t) := ¡ u1 (t), . . . , un (t) of u(x, t) satisfying (5) in [0, τ ]. Then there exists C > 0, depending only on the C 4,1 ([0, 1]×[0, τ ])–norm of u, such that for h small enough we have the following estimate: max max |u(xk , t) − uk (t)| ≤ Ch1/2 .

t∈[0,τ ] 1≤k≤n

(6)

Proof. Let vk (t) := u(xk , t) and ek (t) := vk (t) − uk (t) for 1 ≤ k ≤ n. Observe that Lemma 1 below implies uk (t) ≥ 0 for any t ≥ 0. Let C0 := max{|vk (t)| : 1 ≤ k ≤ n, 0 ≤ t ≤ τ } and t0 := max{t ∈ [0, τ ] : |ek (s)| ≤ C0 /2 for any s ∈ [0, t]}. We shall prove that (6) is valid in the interval [0, t0 ], from which we shall conclude that t0 = τ holds for h small enough. Let k 6= 1, n. Then there exists a constant C1 > 0 independent of h such that ¡ ¢ ¡ ¡ ¢ ¡ ¢¢ e0k (t) = h−2 ¡ek+1 (t) − 2ek (t) + ek−1 (t)¢ − g¡ vk (t)¢ − g uk (t) + C1 h2 = h−2 ek+1 (t) − 2ek (t) + ek−1 (t) − g 0 ξk (t) ek (t) + C1 h2 holds, where ξk (t) in an intermediate value between vk (t) and uk (t). Furthermore, arguing in a similar way for k = 1, n, we obtain: ¡ ¢ ¡ ¢ e01 (t)/2 = h−2 ¡e2 (t) − e1 (t) − g 0 ξ1 (t) e1¢(t)/2¡+ C1 h¢2 /2, 0 e0k (t) = h−2 ¡ek+1 (t) − 2ek (t)¢ + ek−1 + C1 h2 , (2 ≤ k < n) (7) ¡ (t) ¢− g ξk (t) ek (t) 0 −2 0 2 en (t)/2 = h en−1 (t) − en (t) − g ξn (t) en (t)/2 + C1 h /2. Pn−1 Let E(t) := (e1 (t), . . . , en (t)) and N (t) := e21 (t)/2 + k=2 e2k (t) + e2n (t)/2. Multiplying the k-th equality of (7) by ek (t) for 1 ≤ k ≤ n and adding up we have: n−1 ³ ´ X N 0 (t) = 2h−2 E(t)t AE(t) + 2C1 h2 e1 (t)/2 + ek (t) + en (t)/2 , k=2

where A ∈ Zn×n is a suitable negative semidefinite symmetric n × n matrix (the opposite of the stiffness matrix). Therefore, taking into account the inequalities

E(t)t AE(t) ≤ 0 and ek (t) ≤ (e2k (t) + 1)/2 (1 ≤ k ≤ n), we obtain N 0 (t) ≤ C1 h2 N (t) + C1 h. Integrating both members of this inequality we have: Z t Z t N (t) ≤ C1 h2 N (s)ds + C1 th ≤ C1 h2 N (s)ds + C1 T h 0

0

for any t ∈ [0, t0 ]. Therefore, Gronwall’s Lemma (see e.g. [9, §1.2.1]) yields: 2

N (t) ≤ C1 T heC1 T h ≤ C1 T eT C1 h, for any t ∈ [0, t0 ]. Hence, from the definition of N (t) we easily deduce the estimate e2k (t) ≤ 2C1 T eT C1 h for any t ∈ [0, t0 ] and any 1 ≤ k ≤ n. Letting C := (2C1 T )1/2 eT C1 /2 we conclude that |u(xk , t) − uk (t)| ≤ Ch1/2 holds for any 1 ≤ k ≤ n and any t ∈ [0, t0 ]. Furthermore, combining this estimate with the definition of t0 shows that t0 = τ must hold for h small enough, because otherwise the maximality of t0 would be contradicted. This finishes the proof of the theorem. u t Now we analyze the asymptotic behaviour of the solutions of (5). For this purpose, we are going to analyze the role of the stationary solutions of (5), which are the positive solutions of the polynomial system (4). We shall make use of the following discrete maximum principle: Lemma 1. Let U be a solution of (5) with initial condition U (0) = U0 ∈ (R≥0 )n , and let τ ∈ (R>0 ∪ {∞}) be the supremum of the set of t ∈ R>0 for which U is well–defined in [0, t). Then U (t) ∈ (R≥0 )n for any t ∈ [0, τ ). Combining Lemma 1 with e.g. [22, Theorem 1] we conclude that the set of solutions of (5) with positive initial condition is (topologically equivalent to) a dynamical system over (R≥0 )n . Following [4], let Φh : (R≥0 )n → R be the following function: (0)

Φh (U (0) ) := −M (U (0) ) + G(U1 ) +

n−1 X

(0)

2G(Uk ) + G(Un(0) ) − 2αh−1 Un(0) ,

k=2

Pn−1 where M (U ) := 2 k=1 Uk+1 Uk − (U12 + 2 k=2 Uk2 + Un2 ) and G0 (x) := g(x). It is easy to see that Φh is a Liapunov functional for the ¡dynamical sys¡ ¢ tem over (R≥0 )n defined by (5), i.e. Φ0h (u(0) ) := limt→0+ (1/t) Φh φt (u(0) ) − ¢ Φh (u(0) ) ≤ 0 for any u(0) ∈ (R≥0 )n , where φt is the solution of (5) passing through u(0) when t = 0. Furthermore, we have that Φ0h (u(0) ) = 0 holds©if and only if u(0) represents a stationary solution of (5). Hence, defining E := u(0) ∈ ª (R≥0 )n : Φ0h (u(0) ) = 0 , we have that E is invariant under the action of the dynamical system over (R≥0 )n defined by (5). Therefore, from e.g. [9, Theorem 4.3.4] we conclude every solution of (5), with positive initial condition and bounded image, converges to a stationary solution of (5). As a consequence, we see the relevance of the consideration of the set of stationary solutions in order to describe the dynamics of the set of solutions of (5). Pn−1

3

Conditioning and Complexity of our Family of Systems.

In this section we are going to analyze the set of positive solutions of the following polynomial equation system:  ¡ ¢  0 = h−1 U¡1 − U2 + hg(U1 )/2,¢ 0 = −h−1 (2 ≤ k ≤ n − 1) (8) ¡ Uk+1 − 2U ¢ k + Uk−1 + hg(Uk ),  0 = h−1 Un − Un−1 − α + hg(Un )/2, where g is a nonlinear polynomial of Q[T ] with g(0) = g 0 (0) = 0, g 0 (x) > 0 and g 00 (x) ≥ 0 for any x > 0, α ∈ Q>0 and h := (n − 1)−1 . Let us recall that the positive solutions of (8) represent the stationary solutions of the initial value problem (5) of Section 2. The main result of this section asserts that (8) has only one positive solution u∗ ∈ (R≥0 )n , which is well–conditioned (see Theorems 2 and 3 below). This result will allow us to exhibit an algorithm which computes an¡ ε–approximation of u∗ with O(n3 M ) floating point operations, where ¢ M := O log | log(εn)| . Hence, this algorithm significantly improves and generalizes the approach of [6], and exponentially improves the standard homotopy continuation methods, which applied to (8) require L deg(g)O(n) floating-point operations (see e.g. [3]). First, we are going to show that there exists only one positive real solution of (8). Indeed, following [7], [6], we consider the following deformation of (8):  ¡ ¢  0 = h−1 U¡1 − U2 + hg(U1 )/2, ¢ 0 = −h−1¡ Uk+1 − Uk ¢− T (Uk − Uk−1 ) + hg(Uk ), (2 ≤ k ≤ n − 1) (9)  0 = h−1 T Un − Un−1 − α + hg(Un )/2, Let T, U1 , . . . , Un be indeterminates over Q, let U := (U1 , . . . , Un ) and let F : Rn+1 → Rn denote the polynomial mapping defined by the right–hand–side members of (9). From the first n − 1 equations of (9) we easily see that, for given positive values T = t, U1 = u1 , the (positive) values of U2 , . . . , Un are uniquely determined. Therefore, letting U1 and T vary, we may consider U2 , . . . , Un as functions of U1 and T , which are indeed recursively defined as follows: U2 (t, u1 ) := u1 + h2 g(u1 )/2, ¡ ¢ ¡ ¢ U3 (t, u1 ) := U2 (t, u1 ) + t U2 (t, u1 ) − u1 + h2 g U¡2 (t, u1 ) , ¢ (10) Uk+1 (t, u1 ) := Uk (t, u1 ) + t(Uk − Uk−1 )(t, u1 ) + h2 g Uk (t, u1 ) for k ≥ 3. Remark 1. For any u1 > 0 we have: ¡ ¢ Pk (i) (Uk − Uk−1 )(t, u1 ) = h2 tk−2 g(u1 )/2 + j=3 tk−j g(Uj−1 (t, u1 )) > 0 and Uk (t, u1 ) > 0 for 2 ≤ k ≤ n. (ii) (∂Uk /∂U1 − ∂Uk−1 /∂U1 )(t, u1 ) > 0 and (∂Uk /∂U1 )(t, u1 ) > 0 for 2 ≤ k ≤ n. (iii) (∂Uk /∂T − ∂Uk−1 /∂T )(t, u1 ) ≥ 0 and (∂Uk /∂T )(t, u1 ) ≥ 0 for 2 ≤ k ≤ n. Proof. Let k = 2. Then, from (10), we have U2 (t, u1 ) − u1 = h2 g(u1 )/2, (∂U2 /∂U1 )(t, u1 ) = 1 + h2 g 0 (u1 )/2, (∂U2 /∂T )(t, u1 ) = (∂U1 /∂T )(t, u1 ) = 0,

from which we immediately deduce (i), (ii) and (iii) for k = 2. Now, arguing inductively, suppose our statement true for a given k ≥ 2. From (10) we have: (Uk+1 − Uk )(t, u1 ) ∂Uk k+1 ( ∂U ∂U1 − ∂U1 )(t, u1 ) ∂Uk+1 k ( ∂T − ∂U ∂T )(t, u1 )

= t(Uk − Uk−1 )(t, u1 ) + h2 g(Uk (t, u1 )), ∂Uk−1 ∂Uk 2 0 k = t( ∂U ∂U1 − ∂U1 )(t, u1 ) + h g (Uk (t, u1 )) ∂U1 (t, u1 ), ∂Uk−1 ∂Uk = (Uk − Uk−1 )(t, u1 ) + t( ∂T − ∂T )(t, u1 )+ k +h2 g 0 (Uk (t, u1 )) ∂U ∂T (t, u1 ).

Combining these identities with the inductive hypotheses, we easily conclude that (i), (ii) and (ii) hold for k + 1. u t Let Q : [0, 1] × R → R be the polynomial mapping defined by: ¡ ¢ Q(T, U1 ) := T h−1 Un (T, U1 ) − Un−1 (T, U1 ) − α + hg(Un (T, U1 ))/2.

(11)

Observe that Q(T, U1 ) = 0 represents the minimal equation satisfied by the coordinates (t, u1 ) of any (complex) solution of the polynomial equation system F (T, U ) = 0 of (9). Therefore, for fixed t ∈ [0, 1], the positive roots of Q(t, U1 ) are the values of U1 we want to obtain. From Remark 1 we see that Q(t, U1 ) is a strictly increasing function in R≥0 for any t ∈ [0, 1]. Therefore, taking into account that Q(t, 0) < 0 holds and Q(t, U1 ) has a positive leading coefficient, we conclude that there exists a unique positive solution of the system F (t, U ) = 0, i.e. we have shown the following result: Theorem 2. For any t ∈ [0, 1], (9) has only one positive solution. 3.1

An Estimate on the Condition Number of the Positive Solution of (8). ¡ ¢ Let us denote by u1 (t), . . . , un (t) the only positive solution of the polynomial system F (t, U ) = 0 for any t ∈ [0, 1], and observe that in fact it belongs to (R>0 )n . Therefore, we have implicitly defined an analytic function ϕ : [0, 1] → Rn ¡ ¢ by ϕ(t) := u1 (t), . . . , un (t) . Our intention is to analyze the conditioning of approximating the value ϕ(1) by a continuation homotopy method. Following e.g. [3], the condition number of approximating ϕ(t) is given by ° ¡ ¢−1 ° ° ¡ ¢° ° °(∂F/∂T ) t, ϕ(t) ° , kϕ0 (t)k∞ ≤ °(∂F/∂U ) t, ϕ(t) ∞ ∞ t where k · k∞ denotes the standard infinite norm and denotes transposition. ¡ ¢ Let us fix t ∈ [0, 1] and let as above ϕ(t) := u1 (t), . . . , un (t) . In order to ° ° ° ¡ ¢−1 ° ° and °(∂F/∂T )(t, ϕ(t))° , we are going to find estimate °(∂F/∂U ) t, ϕ(t) ∞ ∞ a suitable lower bound for u1 (t) and a suitable upper bound for un (t). For this purpose we have the following technical result, which is a critical point in our estimate on the lower bound of u1 (t) for any t ∈ [0, 1].

Lemma 2. For any t ∈ [0, 1], let (u1 (t), . . . , un (t)) be the positive solution of F (t, U ) = 0. For 2 ≤ k ≤ n we have (i) uk (t) − uk−1 (t) ≤ 2hα. (ii) uk (t) ∈ [u1 (t), u1 (t) + 2α]. Proof. Remark 1 and the condition g 0 (x) > 0 for x > 0 imply g(u1 (t)) < · · · < g(un (t)). On the ¡other hand, combining Remark 1 (i) with the last equation of (9) ¢ we obtain α = h tn−1 g(u1 (t))/2 + tn−2 g(u2 (t)) + · · · + tg(un−1 (t)) + g(un (t))/2 . Therefore we have: u2 (t) − u1 (t) = h2 g(u1 (t))/2 ≤ h2 g(un (t))/2 ≤ 2hα ¡ ¢ uk (t) − uk−1 (t) = h2 t¡k−2 g(u1 (t))/2 + tk−3 g(u2 (t)) + · · · + g(uk−1 (t)) ¢ ≤ 2h2 tk−2 g(un−(k−2) (t)) + · · · + tg(un−1 (t)) + g(un (t))/2 ≤ 2hα for any 3 ≤ k ≤ n − 1. Hence, we conclude that (i) holds. (ii) is an easy consequence of (i). u t A Lower Bound for u1 (t). Let as above consider the polynomial mapping Q : [0, 1] × R → R defined by (11). Since Q(t, u1 (t)) = 0 holds for any t ∈ [0, 1], from the Implicit Function Theorem it follows that U1 may be considered as a function of T , and we obtain the following expression for the derivative of U1 (T ) with respect to the parameter T for any fixed value t ∈ [0, 1]: ¡1 ¢ ∂Un−1 ∂Un t ∂Un h 0 ∂U1 h (Un − Un−1 ) + h ( ∂T − ∂T ) + 2 g (Un ) ∂T (t, u1 ) < 0. (t, u1 ) = − ¡ t ∂Un ¢ ∂T ( − ∂Un−1 ) + h g 0 (Un ) ∂Un (t, u1 ) h ∂U1

∂U1

2

∂U1

Then, we have u1 (1) ≤ u1 (t) ≤ . . . ≤ un (t) ≤ un (0) = g −1 (2α/h) for any t ∈ [0, 1], where g −1 denotes the inverse function of g in [0, ∞). Hence, a lower bound for u1 (1) yields a lower bound for u1 (t) for any t ∈ [0,¡1]. Let t = 1 and let uk ¢:= uk (1) for 1 ≤ k ≤ n. Since α = h g(u1 )/2 + g(u2 ) + · · · + g(un−1 ) + g(un )/2 holds we have g(u1 ) ≤ α ≤ g(un ). Therefore, Lemma 2 implies g −1 (α) − 2α ≤ u1 < · · · < un ≤ g −1 (α) + 2α, which yields a nontrivial lower bound for u1 if g −1 (α) − 2α > 0 holds. Now suppose that g −1 (α) − 2α ≤ 0 holds. Then, from the last equation of (8) we see that α may be considered as a function of U1 , according to the following definition: ¡ ¢ α(U1 ) := h−1 Un (1, U1 ) − Un−1 (1, U1 ) + hg(Un (1, U1 ))/2, Then we have ¡ ∂Un ¢ h¡ ∂Un−1 ∂α ∂Un ¢ = h−1 (1, U1 ) − (1, U1 ) + g 0 (Un ) (1, U1 ) > 0 ∂U1 ∂U1 ∂U1 2 ∂U1 on the positive reals, which shows that α(U1 ) is an increasing function of U1 with α(0) = 0, and hence bijective, on the positive reals. For a given value α0 ∈ R>0 ,

we shall denote by u1 (α0 ) the unique value of U1 for which α(U1 ) = α0 holds. Observe that in this sense U1 is an increasing function of α0 . Now, the conditions satisfied by g imply that for ε > 0 sufficiently small we have g −1 (ε) − 2ε > 0 and there exists a positive real α0 for which g(2α0 ) = α0 . Therefore, we may consider the least positive real α b satisfying g(2b α) = α b. Observe that if g −1 (α) − 2α ≤ 0 holds for a given value α > 0, then u1 (α) > u1 (b α/2) > g −1 (b α/2) − α b > 0. In conclusion, we have the following result: Proposition 1. Let C := g −1 (e α) − 2e α, with α e := α for g −1 (α) − 2α > 0 and α e=α b/2 otherwise. Then we have u1 > C. An Estimate on the Condition Number of Approximating ϕ(t). Let us fix t ∈ [0, 1]. In order to estimate the condition number of approximating ϕ(t) := (u1 (t), . . . , un (t)), we observe that the Jacobian matrix ∂F (t, U )/∂U of F (t, U ) is tridiagonal with the following expression:  −1  h +hg 0 (U1 )/2 −h−1   .  ∂F (t,U )  −h−1 t h−1 (1+t)+hg 0 (U2 ) . .  . :=  .. .. ∂U −1   . . −h −1 −1 0 −h t h t+hg (Un )/2 Following [19], for a given real A := (aij )1≤i,j≤n wePhave the © n × n matrix ª estimate kA−1 k∞ ≤ max1≤i≤n |aii |−1 (1 − µi )−1 , with µi := |aii |−1 j6=i |aij | ¡ ¢ for 1 ≤ i ≤ n. In the case of the matrix (∂F (t, U )/∂U ) ϕ(t) , we have the inequality |akk |−1 (1 − µk )−1 ≤ 2(n − 1)/g 0 (uk (t)) for 1 ≤ k ≤ n. Hence, from the lower bound of Proposition 1 we deduce °¡ ¢ ¡ ¢° ° ∂F (t, U )/∂U −1 t, ϕ(t) °





2 hg 0 (u1 (t))



2 hg 0 (C)

.

(12)

Now we estimate k(∂F/∂T )(t, ϕ(t))k∞ for any t ∈ [0, 1]. For this purpose, ¡ ¢ ¡ ¢t let us observe that (∂F/∂T ) t, ϕ(t) = h−1 0, u2 (t) − u1 (t), . . . , un (t) − un−1 (t) holds. From Lemma 2 we obtain ° ° °(∂F/∂T )(t, ϕ(t))° ≤ 2α. (13) ∞ Combining (12) and (13) we obtain the main result of this section: Theorem 3. The condition number of approximating the only positive solution of F (t, X) = 0 satisfies the estimate κ ≤ 4(n − 1)α/g 0 (C) for any t ∈ [0, 1]. 3.2

An Algorithm Computing the Positive Solution of (8).

As a consequence of the well–conditioning of the positive solution of the system F (t, U ) = 0 of (9) for any t ∈ [0, 1], we shall exhibit an efficient algorithm which computes the only positive solution ϕ(1) of (8). This algorithm is a Newton–Euler

continuation method (see e.g. [20]). For this purpose, let us fix 0 < εb < α/2 and let us introduce for any η ∈ R the polynomial mapping Fη : [0, 1] × Rn → Rn defined in the following way: Fη (T, U ) := F (T, U ) − (0, . . . , 0, η)t . With the same arguments as in the proof of Theorem 2 we conclude that Fη (t, U ) = 0 has only one positive solution for any t ∈ [0, 1] and any η ∈ R with |η| ≤ εˆ. Let f (T ) := −2T 3 + 3T 2 . We have that f (0) = 0, f (1) = 1 and f ([−1/4, 5/4]) = [0, 1]. Then we see that the semi–algebraic subset of R × (R≥0 )n defined by the following system of equalities and inequalities: ¡ ¢ (0, . . . , 0, −b ε)t ≤ F f (T ), U ≤ (0, . . . , 0, εb)t , −1/4 ≤ T ≤ 5/4, is a compact neighborhood of the real algebraic curve F (T, U ) = 0, 0 ≤ T ≤ 1. Observe that this semi–algebraic set may also be defined ¡as the set of points¢ ¡ ¢ t, ϕ(η, t) with t ∈ [−1/4, 5/4] and |η| ≤ εb, where ϕ(η, t) := u1,η (t), . . . , un,η (t) denotes the positive solution of Fη (f (T ), U ) = 0. In order to estimate the complexity of the Newton–Euler method which computes the positive solution of (8), we need an upper bound for un,η (t) and a lower bound for u1,η (t), for any t ∈ [−1/4, 5/4] and any η ∈ [−b ε, εb]. For this purpose, we follow the approach of Section 3.1. Since f (t) ∈ [0, 1] holds for any t ∈ [−1/4, 5/4], we have Cη ≤ u1,η (1) ≤ u1,η (t) ≤ . . . ≤ un,η (t) ≤ un,η (0) = g −1 (2(α + η)/h) ≤ g −1 (4α/h). Therefore, using the estimates of Section 3.1 we conclude that the following estimate holds: °¡ ¢° ¡ ¢ ¢ ¡ ° ∂Fη f (T ), U /∂U −1 t, ϕ(η, t) °





2 e hg 0 (C)

=: β,

e := inf{Cη : η ∈ [−b where C ε, εb]}. Let us remark that our choice of εb implies e C > 0. °¡ ¡ ¢ ¢¡ ¢° We also need an upper bound on ° ∂ 2 Fη f (T ), U /∂U 2 t, ϕ(η, t) °∞ . For this purpose, we have to estimate the norm of the Hessian matrix corresponding to each coordinate¡ of Fη , which ¢ is in turn reduced to estimate the quantity max1≤k≤n {hg 00 (Uk f (t), u1,η (t) )} for any t ∈ [−1/4, 5/4] and any η ∈ [−b ε, εb]. We have the estimate °¡ 2 ¡ ¢ ¢¡ ¢° ° ∂ Fη f (T ),U /∂U 2 t,ϕ(η, t) ° ≤ hγ(n − 1) = γ, ∞ for n ≥ n0 (g, α), where γ and n0 (g, α) are constants independent of n that can be explicitly estimated in terms of the degree and height of g (from e.g. [15] it is easy to obtain such an explicit estimate). Finally, we have ¡ ¢¡ ¢ k ∂Fη (f (T ), U )/∂T t, ϕ(η, t) k∞ ≤ 4α =: δ. Then, applying e.g. [20, 10.4.3], we see that there exists N ≤ 4β 2 γ δ ≤ e 2 )(n − 1)2 = O(n2 ) such that the following holds: (64αγ/(g 0 (C))

If u(0) := ϕ(0) denotes the positive solution of the polynomial system F (0, U ) = 0, and 0 = t0 < t1 < · · · < tN = 1 is a uniform partition of the interval [0, 1], then the iteration ¡ ¢−1 (tk , u(k) )F (tk , u(k) ), (0 ≤ k ≤ N − 1) u(k+1) = u(k) − ∂F (T, U )/∂U yields an attraction point of the standard Newton iteration associated to the system F (1, U ) = 0. Let us remark that, taking into account that the Jacobian matrix (∂F (T, U )/∂U )(tk , u(k) ) is tridiagonal, we conclude that each step of this iteration requires O(L + n) floating-point operations. From [20, 10.4.2–3] we conclude that the vector u(N +k) , obtained from the vector u(N ) above after k steps of the iteration ³¡ ´−1 ¢ u(k+1) = u(k) − ∂F (1, U )/∂U (u(k) ) F (1, u(k) ), (k ≥ N ) satisfies the estimate ku(N +k) − ϕ(1)k∞ ≤ 2−k (2βγ)−1 . Furthermore, combining k−2 this estimate with [20, 10.2.2] we see that ku(N +k) − ϕ(1)k∞ ≤ 2−2 (4βγ)−1 ≤ (k−2) +3) 0 e −1 2−(2 g (C)γ (n − 1)−1 holds for k ≥ 2. Therefore, in order to obtain an ε–approximation of ϕ(1), we have to perform O(M ) steps of the second iteration, e −1 γ)|. Summarizing, we have the following result: with M := log | log(εn(g 0 (C)) Theorem 4. There exists an algorithm which computes ¡ ¢ an ε–approximation of the positive solution of (8) with O (L + n)(n2 + M ) floating-point operations, e −1 γ)|. where M := log | log(εn(g 0 (C))

References [1] E.L. Allgower and K. Georg. Numerical continuation methods: An Introduction, volume 13 of Series in Computational Mathematics. Springer Verlag, Heidelberg, 1990. [2] C. Bandle and H. Brunner. Blow–up in diffusion equations: A survey. Journal of Computational and Applied Mathematics, 97(1–2):3–22, 1998. [3] L. Blum, F. Cucker, M. Shub, and S. Smale. Complexity and Real Computation. Springer, New York Berlin Heidelberg, 1998. [4] J. Fernandez Bonder and J. Rossi. Blow-up vs. spurious steady solutions. Proceedings of the American Mathematical Society, 129(1):139–144, 2001. [5] M. Chipot, M. Fila, and P. Quittner. Stationary solutions, blow up and convergence to stationary solutions for semilinear parabolic equations with nonlinear boundary conditions. Acta Mathematica Universitatis Comenianae, 60(1):35–103, 1991. [6] M. De Leo, E. Dratman, and G. Matera. Numeric vs. symbolic homotopy algorithms in polynomial equation solving: A case study. To appear in J. Complexity, 2004. [7] E. Dratman, and G. Matera. Deformation techniques for counting the real solutions of specific polynomial equation systems. P. D’Argenio, G. Matera (Eds.), Proceedings WAIT’02, Santa Fe, Argentina, September 2002, Vol. 31 of Anales JAIIO, SADIO, Buenos Aires, 2002, pp. 42–52.

[8] R. Ferreira, P. Groisman, and J.D. Rossi. Numerical blow–up for a nonlinear problem with a nonlinear boundary condition. Mathematical models and methods in applied sciences, 12(4):461–484, 2002. [9] D. Henry. Geometric theory of semilinear parabolic equations, volume 840 of Lecture Notes in Mathematics. Springer, Berlin Heideilberg, 1981. [10] J. Heintz, G. Jer´ onimo, J. Sabia, J. San Mart´ın and P. Solern´ o. On the multihomogeneous B´ezout theorem. Manuscript Univ. Buenos Aires (2002). [11] B. Huber, B. Sturmfels, A polyhedral method for solving sparse polynomial systems, Math. Comp. 64 (112) (1995) 1541–1555. [12] H.A. Levine. The role of critical exponents in blow up theorems. SIAM Reviews, 32:262–288, 1990. [13] T. Li. Numerical solution of multivariate polynomial systems by homotopy continuation methods. Acta Numer. 6 (1997) 399–436. [14] T. Li, X. Wang. Solving real polynomial systems with real homotopies. Math. Comp. 60 (1993) 669–680. [15] M. Mignotte. Mathematics for Computer Algebra. Springer, Berlin Heidelberg New York, 1992. [16] A. Morgan. Solving polynomial systems using continuation for engineering and scientific problems. Prentice–Hall, Englewood Cliffs, N.J., 1987. [17] J.J. Mor´e. A collection of nonlinear model problems. In E.L. Allgower and K. Georg, editors, Computational solutions of nonlinear systems of equations, pages 723–762. American Mathematical Society, Providence, RI, 1990. [18] A. Morgan, A. Sommese and C. Wampler. A generic product decomposition formula for B´ezout numbers. SIAM J. Numer. Anal. 32 (1995) 1308–1325. [19] R. Nabben. Two-sided bounds on the inverses of diagonally dominant tridiagonal matrices. Linear Algebra Application, 287:289–305, 1999. [20] J. Ortega and W.C. Rheinboldt. Iterative solutions of nonlinear equations in several variables. Academic Press, New York, 1970. [21] C.V. Pao. Nonlinear parabolic and elliptic equations. Plenum Press, 1992. [22] L. Perko. Differential Equations and Dynamical Systems. Springer-Verlag, New York, 1996. [23] A.A. Samarskii, V.A. Galaktionov, S.P. Kurdyumov, and A.P. Mikhailov. Blow–up in quasilinear parabolic equations, volume 19 of de Gruyter Expositions in Mathematics. Walter de Gruyter, Berlin, 1995. [24] A. Sommese, J. Verschelde and C. Wampler. Numerical decomposition of the solution sets of polynomial systems into irreducible components. SIAM J. Numer. Anal. 38 (6) (2001) 2022–2046. [25] A. Sommese and C. Wampler. Numerical algebraic geometry. In J. Renegar, M. Shub, and S. Smale, editors, The Mathematics of Numerical Analysis: 1995 AMS–SIAM Summer Seminar in Applied Mathematics, July 17–August 11, 1995, Park City, Utah, volume 32 of Lecture in Applied Mathematics, pages 749–763, Providence, RI, 1996. American Mathematical Society.

On the numerical solution of certain nonlinear systems ...

systems arising in semilinear parabolic PDEs⋆. M. De Leo1, E. Dratman2, ... Gutiérrez 1150 (1613) Los Polvorines, Buenos Aires, Argentina. Member of the.

200KB Sizes 1 Downloads 316 Views

Recommend Documents

Approximation of the solution of certain nonlinear ODEs ...
polynomial system solving, condition number, complexity. 2000 MSC: 65H10 ... in [0,l],. (1). Email address: [email protected] (Ezequiel Dratman) .... the comparison between the stationary solutions of (2) and (4). In the second part of the ...

On the Inconsistency of Certain Axioms on Solution ...
moving from the extensive form to the normal form, all strategically relevant information is preserved. It is worth remarking that if (Al) is weakened to require only that two extensive form games have the same solution if their agent-normal forms (s

On the Inconsistency of Certain Axioms on Solution ...
of California at ... At each of a player's information sets, a strategy for that player specifies ... Historically, this encouraged the view that while information is lost in.

NUMERICAL SOLUTION OF BOUSSINESQ SYSTEMS ...
formed in Section 4 on larger spatial intervals and with special initial data that yield nearly one-way .... Application of a 'cleaning' procedure to KdV-KdV systems.

Numerical solution of Boussinesq systems of the Bona ...
Mar 26, 2009 - (cs) data for the solitary waves of the Bona-Smith systems with analogous .... linear mapping Rh : H1 → Sh by aN (Rhv, χ) = aN (v, χ), ∀χ ∈ Sh, ...

Numerical solution of Boussinesq systems of the Bona ...
Appendix of the paper: “Numerical solution of Boussinesq systems of the Bona-Smith family”. D. C. Antonopoulosa, V. A. Dougalisa,b,1and D. E. Mitsotakis?? aDepartment of Mathematics, University of Athens, 15784 Zographou, Greece. bInstitute of Ap

Numerical simulation of nonlinear dynamical systems ...
May 3, 2007 - integration of ordinary, random, and stochastic differential equations. One of ...... 1(yn), v2(yn) and the d × d matrices Bi(yn) are defined by the.

Numerical solution of KdV-KdV systems of Boussinesq ...
Jun 13, 2006 - These are found to be generalized solitary waves, which are symmetric about their crest and which decay to small amplitude periodic structures as the spatial variable becomes large. Key words: Boussinesq systems, KdV-KdV system, genera

Efficient approximation of the solution of certain ...
May 24, 2011 - ing an ε-approximation of such a solution by means of a homotopy contin- .... the point of view of its solution by the so-called robust universal ...

On Global Controllability of Affine Nonlinear Systems with ... - CiteSeerX
We may call the function g1(x)f2(x) − g2(x)f1(x) as the criterion function for global ..... Here, we make a curve γ2 connecting (n − 1)-time smoothly γ1 and the ..... Control Conference, South China University of Technology Press, 2005, pp.

On Global Controllability of Affine Nonlinear Systems with ... - CiteSeerX
We may call the function g1(x)f2(x) − g2(x)f1(x) as the criterion function for global controllability of the system (2.2) ..... one side of the straight-line which passes through the point x0 with direction g(x0. ); ..... bridge: MIT Press, 1973. 1

Numerical solution to the optimal feedback control of ... - Springer Link
Received: 6 April 2005 / Accepted: 6 December 2006 / Published online: 11 ... of the continuous casting process in the secondary cooling zone with water spray control ... Academy of Mathematics and System Sciences, Academia Sinica, Beijing 100080, ..

On the Solution of Linear Recurrence Equations
In Theorem 1 we extend the domain of Equation 1 to the real line. ... We use this transform to extend the domain of Equation 2 to the two-dimensional plane.

The effect of mathematics anxiety on the processing of numerical ...
The effect of mathematics anxiety on the processing of numerical magnitude.pdf. The effect of mathematics anxiety on the processing of numerical magnitude.pdf.

The effect of mathematics anxiety on the processing of numerical ...
The effect of mathematics anxiety on the processing of numerical magnitude.pdf. The effect of mathematics anxiety on the processing of numerical magnitude.pdf.

unit 7 numerical solution of ordinary differential equations
In the previous two units, you have seen how a complicated or tabulated function can be replaced by an approximating polynomial so that the fundamental operations of calculus viz., differentiation and integration can be performed more easily. In this

Studying Nonlinear Dynamical Systems on a Reconfigurable ... - Sites
So, the analog de- signer must depart from the traditional linear design paradigm, ..... [4] B.P. Lathi, Modern Digital and Analog Communication Systems, Oxford.

Numerical solution to the optimal birth feedback ... - Semantic Scholar
Published online 23 May 2005 in Wiley InterScience ... 4 Graduate School of the Chinese Academy of Sciences, Beijing 100049, People's ... the degree of discretization and parameterization is very high, the work of computation stands out and ...

Numerical solution to the optimal birth feedback ... - Semantic Scholar
May 23, 2005 - of a population dynamics: viscosity solution approach ...... Seven different arbitrarily chosen control bi and the optimal feedback control bn.

Download Numerical Methods for Nonlinear Estimating ...
More entrenched forms of nonlinearity often require intensive numerical methods to ... examples from practical applications are included to illustrate the problems and ... Kung-Yee Liang, Scott L. Zeger: Analysis of Longitudinal Data 2/e 26. J.K..

Indexing certain physiological parameters on drought management
S. VINCENT, R. RETHINARAJA AND S. RAJARATHINAM. Coconut Research Station, Veppankulam 614 906, Tamil Nadu. Abstract: The genotype of coconut East Coast Tall was studied for its tolerance of drought through the physiological parameters by using differ

Effect of damping on the nonlinear modal ...
(δ1 ¼0.3%), it is no longer possible to excite the E and F modes, Fig. 8b, they merge presenting the EюF shape, Fig. 7b. The peculiar behavior of the (n:2) internal resonance appears by further increasing the damping level; when c40.03 Ns/m (δ1 4