Efficient approximation of the solution of certain nonlinear reaction–diffusion equation with small absorptionI Ezequiel Dratman Instituto de Ciencias, Universidad Nacional de General Sarmiento, Juan M. Guti´errez 1150 (B1613GSX) Los Polvorines, Buenos Aires, Argentina.

Abstract We study the positive stationary solutions of a standard finite-difference discretization of the semilinear heat equation with nonlinear Neumann boundary conditions. We prove that, if the absorption is small enough, compared with the flux in the boundary, there exists a unique solution of such a discretization, which approximates the unique positive stationary solution of the “continuous” equation. Furthermore, we exhibit an algorithm computing an ε-approximation of such a solution by means of a homotopy continuation method. The cost of our algorithm is linear in the number of nodes involved in the discretization and the logarithm of the number of digits of approximation required. Key words: Two-point boundary-value problem, finite differences, Neumann boundary condition, stationary solution, homotopy continuation, polynomial system solving, condition number, complexity 2000 MSC: 65H10, 65L10, 65L12, 65H20, 65Y20

I The research was partially supported by the following grants: UNGS 30/3084, UNGS 30/1066, CIC (2007–2009) and PIP 11220090100421. Email address: [email protected] (Ezequiel Dratman) URL: http://sites.google.com/site/ezequieldratman (Ezequiel Dratman)

Preprint submitted to Elsevier

May 24, 2011

1. Introduction This article deals with the following semilinear heat equation with Neumann boundary conditions:  ut = uxx − g1 (u) in (0, 1) × [0, T ),    ux (1, t) = αg2 (u(1, t)) in [0, T ), (1) u (0, t) = 0 in [0, T ),  x   u(x, 0) = u0 (x) ≥ 0 in [0, 1], where g1 , g2 ∈ C 3 (R) are analytic functions in x = 0 and α is a positive constant. The nonlinear heat equation models many physical, biological and engineering phenomena, such as heat conduction (see, e.g., [1, §20.3], [2, §1.1]), chemical reactions and combustion (see, e.g., [3, §5.5], [4, §1.7]), growth and migration of populations (see, e.g., [5, Chapter 13], [2, §1.1]), etc. In particular, “power-law” nonlinearities have long been of interest as a tractable prototype of general polynomial nonlinearities (see, e.g., [3, §5.5], [6, Chapter 7], [7], [8], [2, §1.1]). The long-time behavior of the solutions of (1) has been intensively studied (see, e.g., [9], [10], [11], [12], [13], [14], [15], [16] and the references therein). In order to describe the dynamic behavior of the solutions of (1) it is usually necessary to analyze the behavior of the corresponding stationary solutions (see, e.g., [13], [9]), i.e., the positive solutions of the following two-point boundary-value problem:  in (0, 1),  uxx = g1 (u) ux (1) = αg2 (u(1)), (2)  ux (0) = 0. The usual numerical approach to the solution of (1) 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., [17]). This semi-discretization in space leads to the following initial-value problem:  u01 = h22 (u2 − u1 ) − g1 (u1 ),      u0k = h12 (uk+1 − 2uk + uk−1 ) − g1 (uk ), (2 ≤ k ≤ n − 1) (3)  u0n = h22 (un−1 − un ) − g1 (un ) + 2α  h g2 (un ),    uk (0) = u0 (xk ), (1 ≤ k ≤ n) where h := 1/(n−1) and x1 , . . . , xn define a uniform partition of the interval [0, 1]. A similar analysis to that in [18] shows the convergence of the positive 2

solutions of (3) to those of (1) and proves that every bounded solution of (3) tends to a stationary solution of (3), namely to a solution of  0 =   0 =   0 =

2 (u2 − u1 ) − g1 (u1 ), h2 1 (uk+1 − 2uk + uk−1 ) − g1 (uk ), (2 h2 2 (un−1 − un ) − g1 (un ) + 2α h g2 (un ). h2

≤ k ≤ n − 1)

(4) Hence, the dynamic behavior of the positive solutions of (3) is rather determined by the set of solutions (u1 , . . . , un ) ∈ (R>0 )n of (4). Very little is known concerning the study of the stationary solutions of (3) and the comparison between the stationary solutions of (3) and (1). In [13], [18] and [19] there is a complete study of the positive solutions of (4) for the particular case g1 (x) := xp and g2 (x) := xq , i.e., a complete study of the positive solutions of  0 = h22 (u2 − u1 ) − up1 ,   0 = h12 (uk+1 − 2uk + uk−1 ) − upk , (2 ≤ k ≤ n − 1) (5)   p 2 2α q 0 = h2 (un−1 − un ) − un + h un . In [13] it is shown that there are spurious solutions of (4) for q < p < 2q − 1, that is, positive solutions of (4) not converging to any solution of (2) as the mesh size h tends to zero. In [18] and [19] there is a complete study of (4) for p > 2q − 1 and p < q. In these articles it is shown that in such cases there exists exactly one positive real solution. Furthermore, a numeric algorithm solving a given instance of the problem under consideration with nO(1) operations is proposed. In particular, the algorithm of [19] has linear cost in n, that is, this algorithm gives a numerical approximation of the desired solution with O(n) operations. We observe that the family of systems (5) has typically an exponential number O(pn ) of complex solutions ([20]), and hence it is ill conditioned from the point of view of its solution by the so-called robust universal algorithms (cf. [21], [22], [23]). An example of such algorithms is that of general continuation methods (see, e.g., [24]). This shows the need of algorithms specifically designed to compute positive solutions of “structured” systems like (4). Continuation methods aimed at approximating the real solutions of nonlinear systems arising from a discretization of two-point boundary-value problems for second-order ordinary differential equations have been con3

sidered in the literature (see, e.g., [25], [26], [27]). These works are usually concerned with Dirichlet problems involving an equation of the form uxx = f (x, u, ux ) for which the existence and uniqueness of solutions is known. Further, they focus on the existence of a suitable homotopy path rather on the cost of the underlying continuation algorithm. As a consequence, they do not seem to be suitable for the solution of (4). On the other hand, it is worth mentioning the analysis of [28] on the complexity of shooting methods for two-point boundary value problems. Let g1 , g2 ∈ C 3 (R) be analytic functions in x = 0 such that gi (0) = 0, 0 gi (x) > 0, gi00 (x) > 0 and gi000 (x) ≥ 0 for all x > 0 with i = 1, 2. We observe that g1 and g2 are a wide generalization of the monomial functions of system (5). Moreover, we shall assume throughout the paper that the function g := g1 /g2 is strictly decreasing, generalizing thus the relation p < q in (5). In this article we study the existence and uniqueness of the positive solutions of (4), and we obtain numerical approximations of these solutions using homotopy methods. In a forthcoming paper, we shall consider the generalization of the relations q < p < 2q − 1 and 2q − 1 < p in (5). 1.1. Our contributions In the first part of the article we prove that (4) has a unique positive solution, and we obtain upper and lower bounds for this solution independents of h, generalizing the results of [19]. In the second part of the article we exhibit an algorithm which computes an ε-approximation of the positive solution of (4). Such an algorithm is a continuation method that tracks the positive real path determined by the smooth homotopy obtained by considering (4) as a family of systems parametrized by α. Its cost is roughly of n log log ε arithmetic operations, improving thus the exponential cost of general continuation methods. The cost estimate of our algorithm is based on an analysis of the condition number of the corresponding homotopy path, which might be of independent interest. We prove that such a condition number can be bounded by a quantity independent of h := 1/n. This in particular implies that each member of the family of systems under consideration is significantly better conditioned than both an “average” dense system (see, e.g., [29, Chapter 13, Theorem 1]) and an “average” sparse system ([30, Theorem 1]). 1.2. Outline of the paper Section 2 is devoted to determine the number of positive solutions of (4). For this purpose, we prove that the homotopy of systems mentioned

4

above is smooth (Theorem 5). From this result we deduce the existence and uniqueness of the positive solutions of (4). In Section 3 we obtain upper and lower bounds for the coordinates of the positive solution of (4). In Section 4 we obtain estimates on the condition number of the homotopy path considered in the previous section (Theorem 14). Such estimates are applied in Section 5 in order to estimate the cost of the homotopy continuation method for computing the positive solution of (4).

2. Existence and uniqueness of stationary solutions Let U1 , . . . , Un be indeterminates over R. Let g1 and g2 be two functions of class C 3 (R) such that gi (0) = 0, gi0 (x) > 0, gi00 (x) > 0 and gi000 (x) ≥ 0 for all x > 0 with i = 1, 2. As stated in the introduction, we are interested in the positive solutions of (4) for a given positive value of α, that is, in the positive solutions of the nonlinear system  2  0 = −(U2 − U1 ) + h2 g1 (U1 ),   0 = −(Uk+1 − 2Uk + Uk−1 ) + h2 g1 (Uk ), (2 ≤ k ≤ n − 1)   2  0 = −(Un−1 − Un ) h2 g1 (Un ) − hαg2 (Un ), (6) ∗ for a given value α = α > 0, where h := 1/(n − 1). Observe that, as α runs through all possible values in R>0 , one may consider (6) as a family of nonlinear systems parametrized by α, namely,  2  0 = −(U2 − U1 ) + h2 g1 (U1 ),   0 = −(Uk+1 − 2Uk + Uk−1 ) + h2 g1 (Uk ), (2 ≤ k ≤ n − 1)   2  0 = −(Un−1 − Un ) h2 g1 (Un ) − hAg2 (Un ), (7) where A is a new indeterminate. 2.1. Preliminary analysis Let A, U1 , . . . , Un be indeterminates over R, set U := (U1 , . . . , Un ) and denote by F : Rn+1 → Rn the nonlinear map defined by the right-hand side of (7). From the first n − 1 equations of (7) we easily see that, for a given positive value U1 = u1 , the (positive) values of U2 , . . . , Un , A are uniquely

5

determined. Therefore, letting U1 vary, we may consider U2 , . . . , Un , A as functions of U1 , which are indeed recursively defined as follows: U1 (u1 ) := u1 , U2 (u1 ) := u1 +

h2 2 g1 (u1 ),

Uk+1 (u1 ) := 2Uk (u1 ) − Uk−1 (u1 ) + h2 g1 (Uk (u1 )), (2 ≤ k ≤ n − 1),   1 h A(u1 ) := (U − U )(u ) + g (U (u )) /g2 (Un (u1 )). n n−1 1 1 n 1 h 2 (8) Arguing recursively, one deduces the following lemma (cf. [20, Remark 20]). Lemma 1. For any u1 > 0, the following assertions hold:   P g (U (u )) > 0, i. (Uk − Uk−1 )(u1 ) = h2 21 g1 (u1 ) + k−1 j 1 j=2 1   Pk−1 ii. Uk (u1 ) = u1 + h2 k−1 (k − j)g (U (u )) > 0, g (u ) + 1 j 1 1 1 j=2 2 P 0 0 0 )(u1 ) = h2 ( 21 g10 (u1 ) + k−1 iii. (Uk0 − Uk−1 j=2 g1 (Uj (u1 ))Uj (u1 )) > 0, P k−1 0 0 0 iv. Uk0 (u1 ) = 1 + h2 ( k−1 j=2 (k − j)g1 (Uj (u1 ))Uj (u1 )) > 1, 2 g1 (u1 ) + for 2 ≤ k ≤ n. For the proof of the next lema we need the following technical result Remark 2. Let f1 , f2 , f3 ∈ C 2 (R>0 ) be positive functions such that • f100 (x) > 0, • f20 (x) > 0, f30 (x) > 0, • f2 (x) > f3 (x), for all x > 0. Let F : R>0 → R be the function defined by F (x) :=

f1 (f2 (x)) − f1 (f3 (x)) . f2 (x) − f3 (x)

Then F 0 (x) > 0 for all x > 0. Proof. Fix x > 0. By the definition of F we have that   F 0 (x)(f2 (x) − f3 (x)) = f10 (f2 (x))f20 (x) − f10 (f3 (x))f30 (x) −F (x)(f20 (x) − f30 (x))

6

holds. From the Mean Value Theorem, there exists ξ ∈ (f3 (x), f2 (x)) with F (x) = f10 (ξ). Therefore     F 0 (x)(f2 (x) − f3 (x)) = f10 (f2 (x)) − f10 (ξ) f20 (x) + f10 (ξ) − f10 (f3 (x)) f30 (x). Since f10 , f2 and f3 are strictly increasing functions, we conclude that F 0 (x) > 0. Now we prove an important result for the existence and uniqueness of the solutions of (7) Lemma 3. For any u1 > 0, the following assertions hold:   Uk −Uk−1 0 (u1 ) < 0, i. g (U )  1 k 0 −U1 (u1 ) < 0, ii. Ug1k(U k)   Uk −Uk−1 0 iii. (u1 ) ≥ 0, U −U  k 1 0 k) iv. gg11(U (U1 ) (u1 ) > 0, for 2 ≤ k ≤ n. Proof. Let Lj,i : R>0 → R be the function defined by Lj,i (u1 ) :=

g1 (Uj ) − g1 (Ui ) (u1 ), Uj − Ui

where 1 ≤ i < j ≤ n. From Remark 2 and Lemma 1, we deduce that  g (U ) − g (U ) 0 1 j 1 i L0j,i (u1 ) = (u1 ) > 0. Uj − Ui

(9)

By (8) we have 2 −1 U2 − U1 (u1 ) = + L (u ) , 2,1 1 g1 (U2 ) h2 g1 (U2 ) h2 (u1 ) = 1 + L2,1 (u1 ). g1 (U1 ) 2 Combining these identities with (9) we obtain (i), (ii), (iii) and (iv) for k = 2. Now, arguing inductively, suppose that our statement is true for a given k ≥ 2. From (8) we have:   U Uk − U1 −1 k+1 − Uk (u1 ) = 1+ (u1 ) Uk+1 − U1 Uk+1 − Uk  U − U g1 (Uk )h2 −1 −1 k k−1 = 1+ + (u1 ). Uk − U1 Uk − U1 7

Applying the inductive hypotheses we deduce that (iii) holds for k + 1. On the other hand, by Lemma 1 and (8) we have that  g (U ) −1 Uk+1 −Uk 1 k (u1 ) = + Lk+1,k (u1 ) g1 (Uk+1 ) Uk+1 −Uk  U −U −1 −1 k k−1 = + Lk+1,k (u1 ), + h2 g1 (Uk )  Uk+1 −U1 Uk+1 −Uk −1 g1 (Uk ) + Lk+1,k (u1 ) (u1 ) = g1 (Uk+1 ) Uk+1 −U1 Uk+1 −U1  U −U −1 Uk+1 −Uk −1 Uk −U1 k k−1 = + + h2 + Lk+1,k (u1 ), g1 (Uk ) g1 (Uk ) Uk+1 −U1  Uk+1 −U1  g1 (Uk+1 ) (u1 ) = Lk+1,1 (u1 ) + 1 g1 (U1 ) g1 (U1 ) k k − 1 X  g1 (Uj )  = h2 Lk+1,1 + (u1 ) + 1. (k − j) 2 g1 (U1 ) j=2

hold. Combining the inductive hypothesis with (9) and (iii), for k + 1, we conclude that (i), (ii) and (iv) hold for k + 1. 2.2. Existence and uniqueness Let P : (R>0 )2 → R be the nonlinear map defined by P (α, u1 ) := h1 (Un−1 (u1 ) − Un (u1 )) − h2 g1 (Un (u1 )) + αg2 (Un (u1 )).

(10)

Observe that P (A, U1 ) = 0 represents the minimal equation satisfied by the coordinates (α, u1 ) of any (complex) solution of the nonlinear system (7). Therefore, for fixed α > 0, the positive roots of P (α, U1 ) are the values of u1 we want to obtain. Furthermore, from the parametrizations (8) of the coordinates u2 , . . . , un of a given solution (α, u1 , . . . , un ) ∈ (R>0 )n+1 of (7) in terms of u1 , we conclude that the number of positive roots of P (α, U1 ) determines the number of positive solutions of (7) for such a value of α. Therefore, we analyze the existence of positive roots of the function P (α, U1 ) for values α > 0. Let g : R>0 → R>0 be the function defined by g1 (11) g(x) := (x). g2 By Lemma 1(i) we have that P (α, u1 ) = αg2 (Un (u1 )) − h( 12 g1 (u1 ) +

Pn−1

g1 (Uj (u1 )) + 21 g1 (Un (u1 )))   ≥ αg2 (Un (u1 )) − g1 (Un (u1 )) = g2 (Un (u1 )) α − g(Un (u1 )) 8

j=2

holds for any u1 > 0. Suppose that g is surjective. Therefore, there exist u∗1 , u∗∗ 1 > 0 such that g(Un (u∗1 )) = α and g(Un (u∗∗ )) = 2α/h hold. From this choice of u∗1 and u∗∗ 1 1 and the inequality above we deduce   ∗ ∗ ∗ P (α, u1 ) ≥ g2 (Un (u1 )) α − g(Un (u1 )) = 0,   1 ∗∗ ) − U (u∗∗ )) + g (U (u∗∗ )) α − h g(U (u∗∗ )) P (α, u∗∗ ) = (U (u n−1 n 2 n n 1 1 1 1 1 h 2 =

1 ∗∗ h (Un−1 (u1 )

− Un (u∗∗ 1 )) ≤ 0.

Since P (A, U1 ) is a continuous function in (R>0 )2 , from the previous considerations we obtain the following result. Proposition 4. Fix α > 0 and n ∈ N. If the function g of (11) is surjective, then (7) has a positive solution with A = α. In order to establish the uniqueness, we prove that the homotopy path that we obtain by moving the parameter α in R>0 is smooth. For this purpose, we show that the rational function A(U1 ) implicitly defined by the equation P (A, U1 ) = 0 is decreasing. We observe that an explicit expression for this function in terms of U1 is obtained in (8). Theorem 5. Let A(U1 ) be the rational function of (8). If the function g of (11) is decreasing, then the condition A0 (u1 ) < 0 is satisfied for every u1 ∈ R>0 . Proof. Let U1 , U2 , . . . , Un , A be the functions defined in (8). Observe that A can be rewritten as follows: U − U h n n−1 . A = g(Un ) + hg1 (Un ) 2 Taking derivatives with respect to U1 , we have A0 = g 0 (Un )Un0

U − U U − U  h n n−1 n n−1 0 + + g(Un ) . hg1 (Un ) 2 hg1 (Un )

Fix u1 > 0. From Lemma 1 we see that g(Un )(u1 ) is positive. Furthermore, by Lemma 3, we have U − U  n n−1 0 (u1 ) < 0. hg1 (Un )

9

These remarks show that  U − U  h  A  n n−1 A0 (u1 ) < g 0 (Un )Un0 + (u1 ) = g 0 (Un )Un0 (u1 ). (12) hg1 (Un ) 2 g(Un ) From Lemma 1, we deduce that Un0 (u1 ) and A(u1 ) are positive. Combining this affirmation with the monotonicity of g we deduce the statement of the theorem. Now we state and prove the main result of this section. Theorem 6. Let be given α > 0 and n ∈ N. If the function g of (11) is surjective and decreasing, then (6) has a unique positive solution. Proof. Proposition 4 shows that (6) has solutions in (R>0 )n for any α > 0 and any n ∈ N. Therefore, there remains to show the uniqueness assertion. By Theorem 5, the condition A0 (u1 ) < 0 holds for every u1 ∈ R>0 . Arguing by contradiction, assume that there exist two distinct positive solutions (u1 , . . . , un ), (b u1 , . . . , u bn ) ∈ (R>0 )n of (6) for α. This implies that u1 6= u b1 and A(u1 ) = A(b u1 ), where A(U1 ) is defined in (8). But this contradicts the fact that A0 (u1 ) < 0 holds in R>0 , showing thus the theorem. 3. Bounds for the positive solution In this section we obtain bounds for the positive solution of (7). More precisely, we find an interval containing the positive solution of (7) whose endpoints only depend on α. These bounds will allow us to establish an efficient procedure of approximation of this solution. Lemma 7. Let (α, u) ∈ (R>0 )n+1 be a solution of (7) for A = α. Then αg2 (un ) < g1 (un ). Proof. From the last equation of (7) for A = α and Lemma 1(i), we obtain the identity 1  1 αg2 (un ) = h g1 (u1 ) + g1 (u2 ) + · · · + g1 (un−1 ) + g1 (un ) . 2 2 From the previous identity and Lemma 1(i) we immediately deduce the statement of the lemma. From Lemma 7 we obtain the following corollary.

10

Corollary 8. Let (α, u) ∈ (R>0 )n+1 be a solution of (7) for A = α. If the function g of (11) is surjective and strictly decreasing, then un < g −1 (α). Let (α, u) ∈ (R>0 )n+1 be a solution of (7) for A = α. In the following lemma we obtain an upper bound of un in terms of u1 and α. Lemma 9. Let (α, u) ∈ (R>0 )n+1 be a solution of (7) for A = α. If the function g of (11) is surjective and strictly decreasing, then un < eM u1 holds, with M := g10 (g −1 (α)). Proof. Let (α, u) ∈ (R>0 )n+1 be a solution of (7) for A = α. Combining Lemma 1(i) and the Mean Value Theorem, we obtain the following identities   1) uk+1 = uk + h2 g1 (u + g (u ) + · · · + g (u ) 1 2 1 k 2  0  g (ξ )u = uk + h2 1 21 1 + g10 (ξ2 )u2 + · · · + g10 (ξk )uk for 1 ≤ k ≤ (n − 1), where ξi ∈ [0, ui ] for 1 ≤ i ≤ k. Since g10 is an increasing function in R>0 , combining Lemma 1(i) and Corollary 8, we obtain uk+1 ≤ uk + h2 (g10 (u1 )u1 + · · · + g10 (uk )uk ) ≤ (1 + hg10 (g −1 (α)))uk = (1 + hM )uk for 1 ≤ k ≤ (n − 1). Arguing recursively, we deduce that un ≤ (1 + M h)n−1 u1 ≤ eM u1 . This completes the proof. In our next lemma we obtain a lower bound of u1 in terms of α. Lemma 10. Let (α, u) ∈ (R>0 )n+1 be a solution of (7) for A = α. If the function g of (11) is surjective and strictly decreasing, then u1 > g −1 (αC(α)) holds, where C(α) ≥ 1 is a constant such that lim C(α) = 1.

α→+∞

Proof. From Lemma 9 and Lemma 1(i) we deduce the inequalities 11

• g2 (un ) < g2 (eM u1 ),   • g1 (u1 ) < h 12 g1 (u1 ) + g1 (u2 ) + · · · + g1 (un−1 ) + 21 g1 (un ) = αg2 (un ). with M := g10 (g −1 (α)). Combining both inequalities we obtain g1 (u1 ) < αg2 (eM u1 ) = α

g2 (eM u1 ) g2 (u1 ). g2 (u1 )

(13)

Since g2 is an analytic function in x = 0 and g2 (x) 6= 0 for every x > 0, exists k ≥ 1 such that g2 (eM x) lim = ekM . + g (x) x→0 2 Combining this with Corollary 8, we deduce that there exists C(α) > 0 such that g2 (eM u1 ) 1≤ ≤ C(α). g2 (u1 ) Furthermore, we can choose C(α) with lim C(α) = 1.

α→+∞

Combining this remark with (13) we obtain g1 (u1 ) < αC(α)g2 (u1 ), which immediately implies the statement of the lemma. 4. Numerical conditioning Let be given n ∈ N and α∗ > 0. In order to compute the positive solution of (7) for this value of n and A = α∗ , we shall consider (7) as a family of systems parametrized by the values α of A, following the positive real path determined by (7) when A runs through a suitable interval whose endpoints are α∗ and α∗ , where α∗ is a positive constant independent of h to be fixed in Section 5. A critical measure for the complexity of this procedure is the condition number of the path considered, which is essentially determined by the inverse of the Jacobian matrix of (7) with respect to the variables U1 , . . . , Un , and the gradient vector of (7) with respect to the variable A on the path. In this section we prove the invertibility of such a Jacobian matrix, and obtain an explicit form of its inverse. Then we obtain an upper bound on the condition number of the path under consideration. 12

4.1. The Jacobian matrix Let F := F (A, U ) : Rn+1 → Rn be the nonlinear map defined by the right-hand side of (7). In this section we analyze the invertibility of the Jacobian matrix of F with respect to the variables U , namely,   Γ1 −1    −1 . . . . . .  ∂F , J(A, U ) := (A, U ) :=    .. .. ∂U  . . −1  −1 Γn with Γ1 := 1 + 21 h2 g10 (U1 ), Γi := 2 + h2 g10 (Ui ) for 2 ≤ i ≤ n − 1 and Γn := 1 + 21 h2 g10 (Un ) − hAg20 (Un ). We start relating the nonsingularity of the Jacobian matrix J(α, u) with that of the corresponding point in the path determined by (7). Let (α, u) ∈ (R>0 )n+1 be a solution of (7) for A = α. Taking derivatives with respect to U1 in (8) and substituting u1 for U1 we obtain the following tridiagonal system:      Γ1 (u1 ) −1 1 0   0 .. ..   ..  −1   U2 (u1 )  . .    .   =   . .   . . .    . . . 0  . . −1  0 0 Un (u1 ) hg2 (Un (u1 ))A (u1 ) −1 Γn (u1 ) For 1 ≤ k ≤ n − 1, we denote by ∆k := ∆k (A, U ) the kth principal minor of the matrix J(A, U ), that is, the (k×k)-matrix formed by the first k rows and the first k columns of J(A, U ). By the Cramer rule we deduce the identities: hg2 (Un (u1 ))A0 (u1 ) = det (J(α, u)), det (J(α, u))Uk0 (u1 )

(14) 0

= hg2 (Un (u1 ))A (u1 ) det (∆k−1 (α, u)),

(15)

for 2 ≤ k ≤ n. Suppose that α > 0 and that the function g of (11) is decreasing. Then Theorem 5 asserts that A0 (u1 ) < 0 holds. Combining this inequality with (14) we conclude that det (J(α, u)) < 0 holds. Furthermore, by (15), we have Uk0 (u1 ) = det (∆k−1 (α, u))

(2 ≤ k ≤ n).

(16)

Combining Remark 1(iv) and (16) it follows that det (∆k (α, u)) > 0 holds for 1 ≤ k ≤ n−1. As a consequence, we have that all the principal minors of the symmetric matrix ∆n−1 (α, u) are positive. Then the Sylvester criterion shows that ∆n−1 (α, u) is positive definite. These remarks allow us to prove the following result. 13

Theorem 11. Let (α, v) ∈ (R>0 )n+1 be a solution of (7) for A = α. If the function g of (11) is decreasing, then the matrix J(α, u) is invertible with det (J(α, u)) < 0. Furthermore, their (n−1)th principal minor is symmetric and positive definite. Having shown the invertibility of the matrix J(α, u) for every solution (α, u) ∈ (R>0 )n+1 of (7), the next step is to obtain an explicit expression for the corresponding inverse matrices J −1 (α, u). For this purpose, we establish a result on the structure of the matrix J −1 (α, u). Proposition 12. Let (α, u) ∈ (R>0 )n+1 be a solution of (7). If the function g of (11) is decreasing, then the following matrix factorization holds:       J −1 (α, u) =    

1

1 u02

1

1 u03 u02 u03

...

..

..

.

...

1

.

1 u0n u02 u0n

.. . u0n−1 u0n



1 u02

         

1 u03

u02 u03

.. .

.. .

1 u0n

u02 u0n u02

1

1 d(J)



d(J)

..

.

... ...

u0n−1 u0n u0n−1 d(J)

u0n d(J)

     ,    

where d(J) := det (J(α, u)) and u0k := Uk0 (u1 ) for 2 ≤ k ≤ n. Proof. Since J(α, u) is symmetric, invertible, tridiagonal and their (n − 1)th principal minor is positive definite, the proof follows by a similar argument to that of [18, Proposition 25]. 4.2. Upper bounds on the condition number From the explicit expression of the inverse of the Jacobian matrix J(A, U ) on the points of the real path determined by (7), we can finally obtain estimates on the condition number of such a path. Let α∗ > 0 and α∗ > 0 be given constants independent of h. Suppose that the function g of (11) is surjective and decreasing. Then Theorem 6 proves that (7) has a unique positive solution with A = α for every α in the real interval I := I(α∗ , α∗ ) whose endpoints are α∗ and α∗ , which we denote by (u1 (α), U2 (u1 (α)), . . . , Un (u1 (α))). We bound the condition number κ := max{kϕ0 (α)k∞ : α ∈ I}, associated to the function ϕ : I → Rn , ϕ(α) := (u1 (α), U2 (u1 (α)), . . . , Un (u1 (α))). 14

For this purpose, from the Implicit Function Theorem we have

 ∂F

−1 ∂F

kϕ0 (α)k∞ = (α, ϕ(α)) (α, ϕ(α)) ∂U ∂A ∞

∂F

−1 (α, ϕ(α)) . = J (α, ϕ(α)) ∂A ∞  t We observe that (∂F/∂A)(α, ϕ(α)) = 0, . . . , 0, −hg2 (Un (u1 (α))) holds. From Proposition 12 we obtain

hg (U (u (α)))  t

2 n 1

kϕ0 (α)k∞ = 1, U20 (u1 (α)), . . . , Un0 (u1 (α)) . det (J(α, ϕ(α))) ∞ Combining this identity with (14), we conclude that

kϕ0 (α)k∞ =

 t 1

0 0 1, U (u (α)), . . . , U (u (α))

. 2 1 n 1 A0 (u1 (α)) ∞

From Lemma 1, we deduce the following proposition. Proposition 13. Let α∗ > 0 and α∗ > 0 be given constants independent of h. Suppose that the function g of (11) is surjective and decreasing. Then kϕ0 (α)k∞ =

Un0 (u1 (α)) |A0 (u1 (α))|

holds for α ∈ I. Combining Proposition 13 and (12) we conclude that kϕ0 (α)k∞ <

g (Un (u1 (α))) α|g 0 (Un (u1 (α)))|

.

Applying Lemma 10 and Corollary 8 we deduce the following result. Theorem 14. Let α∗ > 0 and α∗ > 0 be given constants independent of h. Suppose that the function g of (11) is surjective and g 0 (x) < 0 holds for all x ∈ R>0 . Then there exists a constant κ1 (α∗ , α∗ ) > 0 independent of h with κ < κ1 (α∗ , α∗ ).

15

5. An efficient numerical algorithm As a consequence of the well conditioning of the positive solutions of (7), we shall exhibit an algorithm computing the positive solution of (7) for A = α∗ . This algorithm is a homotopy continuation method (see, e.g., [31, §10.4], [29, §14.3]) having a cost which is linear in n. There are two different approaches to estimate the cost of our procedure: using Kantorovich–type estimates as in [31, §10.4], and using Smale–type estimates as in [29, §14.3]. We shall use the former, since we are able to control the condition number in suitable neighborhoods of the real paths determined by (7). Furthermore, the latter does not provide significantly better estimates. Let α∗ > 0 be a constant independent of h. Suppose that the following conditions hold: • g is surjective, • g 0 (x) < 0 holds for all x > 0, • g 00 (x) ≥ 0 holds for all x > 0, where g is the function of (11). Then the path defined by the positive solutions of (7) with α ∈ [α∗ , α∗ ] is smooth, and the estimate of Theorem 14 hold. Assume that we are given a suitable approximation u(0) of the positive solution ϕ(α∗ ) of (7) for A = α∗ . In this section we exhibit an algorithm which, on input u(0) , computes an approximation of ϕ(α∗ ). We recall that ϕ denotes the function which maps each α > 0 to the positive solution of (7) for A = α. From Corollary 8 and Lemma 10, we have that the coordinates of the positive solution of (7) tend to zero when α tends to infinity. Therefore, for α large enough, we obtain a suitable approximation of the positive solution (7) for A = α∗ , and we track the positive real path determined by (7) until A = α∗ . In order to deal with a bounded set, we consider the change of variables B := 1/A. Then system (7) for A = α∗ can be rewritten in terms of B as follows:  2  0 = −(U2 − U1 ) + h2 g1 (U1 ),      0 = −(Uk+1 − 2Uk + Uk−1 ) + h2 g1 (Uk ), (2 ≤ k ≤ n − 1),       0 = −(U h2 −1 n−1 − Un ) − hB g2 (Un ) + 2 g1 (Un ) (17) 16

for B = β ∗ , where β ∗ := 1/α∗ . Let 0 < β∗ < β ∗ be a constant independent of h to be determined. Fix β ∈ [β∗ , β ∗ ]. By Corollary 8 it follows that ϕ(α) is an interior point of the compact set Kβ := {u ∈ Rn : kuk∞ ≤ 2g −1 (1/β)}, where ϕ : [β∗ , β ∗ ] → Rn is the function which maps each β ∈ [β∗ , β ∗ ] to the positive solution of (17) for B = β, namely ϕ(β) := (u1 (β), . . . , un (β)) := (u1 (β), U2 (u1 (β)), . . . , Un (u1 (β))). First we prove that the Jacobian matrix Jβ (u) := (∂F /∂U )(β, u) is invertible in a suitable subset of Kβ . Let u ∈ Rn and v ∈ Rn be points with ku − ϕ(β)k∞ < δβ , kv − ϕ(β)k∞ < δβ , where δβ > 0 is a constant to be determined. Note that if δβ ≤ g −1 (1/β) then u ∈ Kβ and v ∈ Kβ . By the Mean Value Theorem, we see that the entries of the diagonal matrix Jβ (u) − Jβ (v) satisfy the estimates (1 ≤ i ≤ n − 1), (Jβ (u) − Jβ (v))ii ≤ 2h2 g100 (2g −1 (1/β))δβ (Jβ (u) − Jβ (v))nn ≤ 2h max{g200 (2g −1 (1/β))/β, g100 (2g −1 (1/β))}δβ . By Theorem 11 and Proposition 12 we have that the matrix Jϕ(β) := Jβ (ϕ(β)) = (∂F /∂U )(β, ϕ(β)) is invertible and −1 (Jϕ(β) )ij =

n−1 X k=max{i,j}

Ui0 (u1 (β))Uj0 (u1 (β)) Ui0 (u1 (β))Uj0 (u1 (β)) + 0 Uk0 (u1 (β))Uk+1 (u1 (β)) Un0 (u1 (β)) det(Jϕ(β) )

holds for 1 ≤ i, j ≤ n. According to Lemma 1, we have Un0 (u1 (β)) ≥ · · · ≥ U20 (u1 (β)) ≥ 1. These remarks show that

−1

Jϕ(β) (Jβ (u) − Jβ (v)) ≤ ∞

 h2 + ≤ ηβ δβ 2 +

Pn−1 j=2

h2 Uj0 (u1 (β)) + hUn0 (u1 (β))  | det(Jϕ(β) )|

 hUn0 (u1 (β))  ≤ 2ηβ δβ 1 + , | det(Jϕ(β) )| 17

(18)

where ηβ := 2 max{g100 (2g −1 (1/β)), g200 (2g −1 (1/β))/β}. Since B = 1/A, from Theorem 5 we deduce that B 0 (u1 ) = −A0 (u1 )/A2 (u1 ) > 0 for all x > 0. Combining this assertion with (14), we obtain the following identity: hUn0 (u1 (β)) U 0 (u1 (β))B 2 (u1 (β)) = n0 . | det(Jϕ(β) )| B (u1 (β))g2 (un (β)) From (12), we have that hUn0 (u1 (β)) U 0 (u1 (β))B 2 (u1 (β)) g(un (β))B(u1 (β)) = n0 ≤ 0 . | det(Jϕ(β) )| B (u1 (β))g2 (un (β)) |g (un (β))|g2 (un (β))

(19)

Combining this inequality with the definition and the monotonicity of g, we deduce |g 0 (un (β))|g2 (un (β)) g(un (β))

g1 (un (β))g20 (un (β)) − g10 (un (β))g2 (un (β)) g1 (un (β))  g 0 (un (β))g2 (un (β))  = g20 (un (β)) 1 − 1 . g1 (un (β))g20 (un (β)) =

Since g1 and g2 are analytic functions in x = 0 and g1 (0) = g2 (0) = 0, there exists r > 0 such that, for all |x| < r, we have that g1 (x) =

∞ X

ck xk , g2 (x) =

∞ X

dk xk ,

k=q

k=p

with cp 6= 0 and dq 6= 0, where p and q are positive integers greater than 1. Hence, the following identity holds:   g10 (x)g2 (x) p lim 1 − =1− . 0 x→0 g1 (x)g2 (x) q Taking into account that un (β) ∈ (0, g −1 (1/β ∗ )] holds for all β ∈ [β∗ , β ∗ ], we conclude |g 0 (un (β))|g2 (un (β)) ≥ g20 (un (β))(1 − ρ∗ ), g(un (β)) where ρ∗ is a constant which depends only on β ∗ . Combining the last inequality with (18) and (19), we obtain

 

−1

Jϕ(β) Jβ (u) − Jβ (v)



 g(un (β))B(u1 (β))  ≤ 2ηβ δβ 1 + 0 |g (un (β))|g2 (un (β))   B(u1 (β)) ≤ 2ηβ δβ 1 + 0 . g2 (un (β))(1 − ρ∗ ) 18

From Lemma 10 and Corollary 8, there exists a constant C(β) ≥ 1 such that

 

−1

Jϕ(β) Jβ (u) − Jβ (v)



 ≤ 2ηβ 1 +

 β δβ . (20) g20 (g −1 (C(β)/β))(1 − ρ∗ )

Furthermore, the following condition holds: lim C(β) = 1.

β→0+

Finally, since g 0 (x) < 0 holds for all x > 0, we have β g20 (g −1 (C(β)/β))(1



ρ∗ )

= ≤

C(β) g20 (g −1 (C(β)/β))g(g −1 (C(β)/β))(1

− ρ∗ )

C(β) . g10 (g −1 (C(β)/β))(1 − ρ∗ )

Combining this inequality with (20), we deduce that

 

−1

Jϕ(β) Jβ (u) − Jβ (v)





2ηβ (θ∗ + 1)C(β) δβ , g10 (g −1 (C(β)/β))(1 − ρ∗ )

(21)

with θ∗ := (1 − ρ∗ )g10 (g −1 (1/β ∗ )). Hence, defining δβ in the following way: δβ := min

n g 0 (g −1 (C(β)/β))(1 − ρ∗ ) 1

8ηβ (θ∗ + 1)C(β)

o , g −1 (1/β) ,

(22)

we obtain

1 (23) ≤ . 4 ∞ In particular, for v = ϕ(β), this bound allows us to consider Jβ (u) as a perturbation of Jϕ(β) . More precisely, by a standard perturbation lemma (see, e.g., [31, Lemma 2.3.2]) we deduce that Jβ (u) is invertible for every u ∈ Bδβ (ϕ(β)) ∩ Kβ and we obtain the following upper bound:

 

−1

Jϕ(β) Jβ (u) − Jβ (v)



Jβ (u)−1 Jϕ(β)



4 ≤ . 3

(24)

In order to describe our method, we need a sufficient condition for the convergence of the standard Newton iteration associated to (7) for any β ∈ [β∗ , β ∗ ]. Arguing as in [31, 10.4.2] we deduce the following remark, which in particular implies that the Newton iteration under consideration converges.

19

Remark 15. Set δ := min{δβ : β ∈ [β∗ , β ∗ ]}. Fix β ∈ [β∗ , β ∗ ] and consider the Newton iteration u(k+1) = u(k) − Jβ (u(k) )−1 F (β, u(k) )

(k ≥ 0),

starting at u(0) ∈ Kβ . If ku(0) − ϕ(β)k∞ < δ, then ku(k) − ϕ(β)k∞ <

δ 3k

holds for k ≥ 0. Now we can describe our homotopy continuation method. Let β0 := β∗ < β1 < · · · < βN := β ∗ be a uniform partition of the interval [β∗ , β ∗ ], with N to be fixed. We define an iteration as follows: u(k+1) = u(k) − Jβk (u(k) )−1 F (βk , u(k) ) (N +k+1)

u

(N +k)

= u

(N +k) −1

− Jβ ∗ (u

)



(0 ≤ k ≤ N − 1),

(25)

(N +k)

(26)

F (β , u

)

(k ≥ 0).

In order to see that the iteration (25)–(26) yields an approximation of the positive solution ϕ(β ∗ ) of (17) for B = β ∗ , it is necessary to obtain a condition assuring that (25) yields an attraction point for the Newton iteration (26). This relies on a suitable choice for N , which we now discuss. By Theorem 14, we have kϕ(βi+1 ) − ϕ(βi )k∞ ≤ max{kϕ0 (β)k∞ : β ∈ [β∗ , β ∗ ]} |βi+1 − βi | β∗ ≤ κ1 , N for 0 ≤ i ≤ N − 1, where κ1 is an upper bound of the condition number independent of h. Thus, for N := d3β ∗ κ1 /δe + 1 = O(1), by the previous estimate we obtain the following inequality: kϕ(βi+1 ) − ϕ(βi )k∞ <

δ 3

(27)

for 0 ≤ i ≤ N − 1. Our next result shows that this implies the desired result. Lemma 16. Set N := d3β ∗ κ1 /δe + 1. Then, for every u(0) with ku(0) − ϕ(β∗ )k∞ < δ, the point u(N ) defined in (25) is an attraction point for the Newton iteration (26).

20

Proof. By hypothesis, we have ku(0) − ϕ(β∗ )k∞ < δ. Arguing inductively, suppose that ku(k) − ϕ(βk )k∞ < δ holds for a given 0 ≤ k < N . By Remark 15 we have that u(k) is an attraction point for the Newton iteration associated to (17) for B = βk . Furthermore, Remark 15 also shows that ku(k+1) − ϕ(βk )k∞ < δ/3 holds. Then ku(k+1) − ϕ(βk+1 )k∞ ≤ ku(k+1) − ϕ(βk )k∞ + kϕ(βk ) − ϕ(βk+1 )k∞ <

1 3δ

+ 31 δ < δ,

where the inequality kϕ(βk+1 ) − ϕ(βk )k∞ < δ/3 follows by (27). This completes the inductive argument and shows in particular that u(N ) is an attraction point for the Newton iteration (26). Next we consider the convergence of (26), starting with a point u(N ) satisfying the condition ku(N ) − ϕ(β ∗ )k∞ < δ ≤ δβ ∗ . Combining this inequality with (22) we deduce that u(N ) ∈ Kα∗ . Furthermore, we see that ku(N +1) −ϕ(β ∗ )k∞ = ku(N ) −Jβ ∗ (u(N ) )−1 F (β ∗ , u(N ) )−ϕ(β ∗ )k∞

 

= Jβ ∗ (u(N ) )−1 Jβ ∗ (u(N ) ) u(N ) −ϕ(β ∗ ) −F (β ∗ , u(N ) )+F (β ∗ , ϕ(β ∗ ))



) −1 ≤ kJβ ∗ (u(N

) Jϕ(β ∗ ) k∞  

−1

Jϕ(β ∗ ) Jβ ∗ (u(N ) ) u(N ) −ϕ(β ∗ ) −F (β ∗ , u(N ) )+F (β ∗ , ϕ(β ∗ ))   ∞ −1 (N ) )−Jβ ∗ (ξ) k∞ k u(N ) −ϕ(β ∗ ) k∞ , ≤ kJβ ∗ (u(N ) )−1 Jϕ(β ∗ ) k∞ kJϕ(β ∗ ) Jβ ∗ (u (28)

where ξ is a point in the segment joining the points u(N ) and ϕ(β ∗ ). Combining (21) and (24) we deduce that ku(N +1) − ϕ(β ∗ )k∞ < <

−1 4 (N ) )−J ∗ (ξ))k δ ∗ ∗ β ∞ β 3 kJϕ(β ∗ ) (Jβ (u 1 ∗ 4c 2 3 δβ ∗ ≤ 3 δβ

holds, with c := (2ηβ ∗ (θ∗ + 1)C(β ∗ ))/(g10 (g −1 (C(β ∗ )/β ∗ ))(1 − ρ∗ )). By an inductive argument we conclude that the iteration (26) is well-defined and converges to the positive solution ϕ(β ∗ ) of (17) for B = β ∗ . Furthermore, we conclude that the point u(N +k) , obtained from the point u(N ) above after k steps of the iteration (26), satisfies the estimate  1 2k 2k ∗ ) , ku(N +k) − ϕ(β ∗ )k∞ ≤ cˆ( 4c δ ≤ c ˆ 3 β 3 with cˆ := 3/4c. Therefore, in order to obtain an ε-approximation of ϕ(β ∗ ), we have to perform log2 log3 (3/4cε) steps of the iteration (26). Summarizing, we have the following result. 21

Lemma 17. Let ε > 0 be given. Then, for every u(N ) ∈ (R>0 )n satisfying the condition ku(N ) − ϕ(β ∗ )k∞ < δ, the iteration (26) is well-defined and the estimate ku(N +k) − ϕ(β ∗ )k∞ < ε holds for k ≥ log2 log3 (3/4cε). Let ε > 0. Assume that we are given u(0) ∈ (R>0 )n such that ku(0) − ϕ(β∗ )k∞ < δ holds. In order to compute an ε-approximation of the positive solution ϕ(β ∗ ) of (17) for B = β ∗ , we perform N iterations of (25) and k0 := dlog2 log3 (3/4cε)e iterations of (26). From Lemmas 16 and 17 we conclude that the output u(N +k0 ) of this procedure satisfies the condition ku(N +k0 ) − ϕ(β ∗ )k∞ < ε. Observe that the Jacobian matrix Jβ (u) is tridiagonal for every β ∈ [β∗ , β ∗ ] and every u ∈ Kβ . Therefore, the solution of a linear system with matrix Jβ (u) can be obtained with O(n) flops. This implies that each iteration of both (25) and (26) requires O(n) flops. In conclusion, we have the following result. Proposition 18. Let ε > 0 and u(0) ∈ (R>0 )n with ku(0) − ϕ(β∗ )k∞ < δ be given, where δ is defined as in Remark 15. Then the output of the iteration (25)–(26) is an ε-approximation of the positive solution ϕ(β ∗ ) of (17) for B = β ∗ . This iteration can be computed with O(N n + k0 n) = O(n log2 log2 (1/ε)) flops. Finally, we exhibit a starting point u(0) ∈ (R>0 )n satisfying the condition of Proposition 18. Let β∗ > 0 be a constant independent of h to be determined. We study the constant δ := min{δβ : β ∈ [β∗ , β ∗ ]}, where δβ := min

n g 0 (g −1 (C(β)/β))(1 − ρ∗ ) 1

8ηβ (θ∗ + 1)C(β)

o , g −1 (1/β) .

Since g1 and g2 are analytic functions in x = 0, in a neighborhood of 0 ∈ Rn , we can rewrite ηβ as follows: ηβ = 2 max{g100 (2g −1 (1/β)), g200 (2g −1 (1/β))g(g −1 (1/β))} = 2 max{S1 (β), S2 (β)}(g −1 (1/β))p−2 , where p is the multiplicity of 0 as a root of g1 and Si is an analytic function in x = 0 such that limβ→0 Si (β) 6= 0 for i = 1, 2. Taking into account that β ∈ (0, β ∗ ] holds, we conclude that there exists a constant η ∗ > 0, which depends only on β ∗ , with ηβ ≤ 2η ∗ (g −1 (1/β))p−2 22

for all β ∈ (0, β ∗ ]. Moreover, with a similar argument we deduce that there exists a constant ϑ∗ > 0, which depends only on β ∗ , such that n  g −1 (C(β)/β) p−1 o ϑ∗ (1 − ρ∗ ) δβ ≥ min , 1 g −1 (1/β). (29) 16η ∗ (θ∗ + 1)C(β) g −1 (1/β) We claim that lim

β→0+

g −1 (C(β)/β) = 1− . g −1 (1/β)

(30)

In fact, since we have C(β) ≥ 1 and g −1 is decreasing, it follows that g −1 (C(β)/β) ≤ 1. g −1 (1/β)

(31)

On the other hand, there exists ξ ∈ (1/β, C(β)/β) with g −1 (C(β)/β) g −1 (1/β)

g −1 (1/β) + (g −1 )0 (ξ)((C(β) − 1)/β) g −1 (1/β) −1 g(g (1/β)) = 1 + 0 −1 (C(β) − 1) g (g (ξ))g −1 (1/β) g(g −1 (1/β)) (C(β) − 1). ≥ 1 + 0 −1 g (g (1/β))g −1 (1/β)

=

(32)

Since g1 and g2 are analytic functions in x = 0 and limβ→0+ C(β) = 1, we see that g(g −1 (1/β)) lim 0 −1 (C(β) − 1) = 0. β→0+ g (g (1/β))g −1 (1/β) Combining (31), (32) and this inequality we inmediately deduce (30). Combining (29) with (30) we conclude that there exists a constant C ∗ ∈ (0, 1], which depends only on β ∗ , with δβ ≥ C ∗ g −1 (1/β). Therefore, δ = min{δβ : β ∈ [β∗ , β ∗ ]} ≥ C ∗ g −1 (1/β∗ ).

(33)

From Corollary 8 and Lemma 10, we have ϕ(β∗ ) ∈ [g −1 (C(β∗ )/β∗ ), g −1 (1/β∗ )]n . Furthermore, by (30), we deduce that 

1−

g −1 (C(β∗ )/β∗ )  −1 g (1/β∗ ) < C ∗ g −1 (1/β∗ ) g −1 (1/β∗ ) 23

(34)

holds for β∗ > 0 small enough. Combining this with (33), we conclude that ku − ϕ(β∗ )k∞ ≤ g −1 (1/β∗ ) − g −1 (C(β∗ )/β∗ ) < δ

(35)

holds for all u ∈ [g −1 (C(β∗ )/β∗ ), g −1 (1/β∗ )]n . Thus, let β∗ < β ∗ satisfy (34). Then, for any u(0) in the hypercube [g −1 (C(β∗ )/β∗ ), g −1 (1/β∗ )]n , the inequality ku(0) − ϕ(β∗ )k∞ < δ holds. Therefore, applying Proposition 18, we obtain our main result. Theorem 19. Let ε > 0 be given. Then we can compute an ε-approximation of the positive solution of (17) for B = β ∗ with O(n log2 log2 (1/ε)) flops. References [1] J. Cannon, The one-dimensional heat equation, Cambridge Univ. Press, Cambridge, 1984. [2] C. Pao, Nonlinear parabolic and elliptic equations, Plenum Press, 1992. [3] J. Bebernes, D. Eberly, Mathematical problems from combustion theory, Vol. 83 of Applied Mathematical Sciences, Springer-Verlag, New York, 1989. [4] P. Grindrod, The theory and applications of reaction-diffusion equations: patterns and waves, Clarendon Press, Oxford, 1996. [5] J. Murray, Mathematical biology. Vol. 1: An introduction, Vol. 17 of Interdisciplinary Applied Mathematics, Springer, New York, 2002. [6] B. Gilding, R. Kersner, Travelling waves in nonlinear diffusionconvection reaction, Birkh¨auser, Berlin, 2004. [7] H. Levine, The role of critical exponents in blow up theorems, SIAM Rev. 32 (1990) 262–288. [8] A. Samarskii, V. Galaktionov, S. Kurdyumov, A. Mikhailov, Blow–up in quasilinear parabolic equations, Vol. 19 of de Gruyter Expositions in Mathematics, Walter de Gruyter, Berlin, 1995. [9] M. Chipot, M. Fila, P. Quittner, Stationary solutions, blow up and convergence to stationary solutions for semilinear parabolic equations with nonlinear boundary conditions, Acta Math. Univ. Comenian. (N.S.) 60 (1) (1991) 35–103. 24

[10] J. Lopez Gomez, V. Marquez, N. Wolanski, Dynamic behaviour of positive solutions to reaction–diffusion problems with nonlinear absorption through the boundary, Rev. Un. Mat. Argentina 38 (1993) 196–209. [11] P. Quittner, On global existence and stationary solutions for two classes of semilinear parabolic problems, Comment. Math. Univ. Carolin. 34 (1) (1993) 105–124. [12] J. Rossi, The blow-up rate for a semilinear parabolic equation with a nonlinear boundary condition, Acta Math. Univ. Comenian. (N.S.) 67 (2) (1998) 343–350. [13] J. Fernandez Bonder, J. Rossi, Blow-up vs. spurious steady solutions, Proc. Amer. Math. Soc. 129 (1) (2001) 139–144. [14] A. Rodr´ıguez Bernal, A. Tajdine, Nonlinear balance for reaction– diffusion equations under nonlinear boundary conditions: Dissipativity and blow–up, J. Differential Equations 169 (2) (2001) 332–372. [15] F. Andreu, J. M. Maz´ on, J. Toledo, J. D. Rossi, Porous medium equation with absorption and a nonlinear boundary condition, Nonlinear Anal. 49 (2002) 541–563. [16] M. Chipot, P. Quittner, Equilibria, connecting orbits and a priori bounds for semilinear parabolic equations with nonlinear boundary conditions, J. Dynam. Differential Equations 16 (2004) 91–138. [17] C. Bandle, H. Brunner, Blowup in diffusion equations: a survey, J. Comput. Appl. Math. 97 (1998) 3–22. [18] E. Dratman, G. Matera, On the solution of the polynomial systems arising in the discretization of certain ODEs, Computing 85 (4) (2009) 301–337. [19] E. Dratman, Approximation of the solution of certain nonlinear ODEs with linear complexity, J. Comput. Appl. Math. 233 (9) (2010) 2339– 2350. [20] M. De Leo, E. Dratman, G. Matera, Numeric vs. symbolic homotopy algorithms in polynomial system solving: A case study, J. Complexity 21 (4) (2005) 502–531. [21] L. Pardo, Universal elimination requires exponential running time, in: A. Montes (Ed.), Computer Algebra and Applications, Proceedings of EACA–2000, Barcelona, Spain, September 2000, 2000, pp. 25–51. 25

[22] D. Castro, M. Giusti, J. Heintz, G. Matera, L. Pardo, The hardness of polynomial equation solving, Found. Comput. Math. 3 (4) (2003) 347–420. [23] E. Dratman, G. Matera, A. Waissbein, Robust algorithms for generalized Pham systems, Comput. Complexity 18 (1) (2009) 105–154. [24] E. Allgower, K. Georg, Numerical continuation methods: An Introduction, Vol. 13 of Series in Computational Mathematics, Springer Verlag, Heidelberg, 1990. [25] E. Allgower, D. Bates, A. Sommese, C. Wampler, Solution of polynomial systems derived from differential equations, Computing 76 (1–2) (2006) 1–10. [26] J. Duvallet, Computation of solutions of two–point boundary value problems by a simplicial homotopy algorithm, in: E. Allgower, K. Georg (Eds.), Computational Solution of Nonlinear Systems of Equations, Vol. 26 of Lectures in Appl. Math., Amer. Math. Soc., Providence, RI, 1990, pp. 135–150. [27] L. T. Watson, Solving finite difference approximations to nonlinear twopoint boundary value problems by a homotopy method, SIAM J. Sci. Stat. Comput. 1 (4) (1980) 467–480. [28] B. Kacewicz, Complexity of nonlinear two-point boundary-value problems, J. Complexity 18 (3) (2002) 702–738. [29] L. Blum, F. Cucker, M. Shub, S. Smale, Complexity and Real Computation, Springer, New York Berlin Heidelberg, 1998. [30] G. Malajovich, J. Rojas, High probability analysis of the condition number of sparse polynomial systems, Theoret. Comput. Sci. 315 (2–3) (2004) 525–555. [31] J. Ortega, W. Rheinboldt, Iterative solutions of nonlinear equations in several variables, Academic Press, New York, 1970.

26

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 ...

224KB Sizes 3 Downloads 421 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 ...

EFFICIENT APPROXIMATION OF THE SOLUTION OF ...
an ε-approximation of such a solution by means of a homotopy continuation .... view of its solution by the so-called robust universal algorithms (cf. [26], [8], [16]).

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.

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.

Efficient Distributed Approximation Algorithms via ...
a distributed algorithm for computing LE lists on a weighted graph with time complexity O(S log n), where S is a graph .... a node is free as long as computation time is polynomial in n. Our focus is on the ...... Given the hierarchical clustering, o

Efficient MAP approximation for dense energy functions - The Robotics ...
Appearing in Proceedings of the 23rd International Con- .... Lagrange multipliers we obtain the free energy as: ..... Conference on Computer Vision, 2005.

Enforceability of Certain Agreements Between the Department of the ...
Sep 26, 2008 - 1491(a)(1) of title 28 of the United States Code, authorizes the Court of Federal. Claims “to render judgment upon any claim against the United ...

TARIFF TREATMENT OF CERTAIN ... - WorldTradeLaw.net
Feb 27, 2015 - RUSSIA - TARIFF TREATMENT OF CERTAIN AGRICULTURAL. AND MANUFACTURING PRODUCTS ... agricultural and manufacturing sectors. On 31 October 2014, the European Union requested ... 1 January 2016 (sixth measure at issue). Secondly, for certa

Approximation of the Invariant Probability Measure of ...
Sep 14, 2007 - Since Pg E 9+the last summand tends to zero as k*~. Therefore we ... which is not a consequence of (ii)' and lim sup,,, Pmgmd Pg. Of course.

Global Solver and Its Efficient Approximation for ...
subspace clustering (LRSC) by providing an exact global solver and its efficient ... There, to avoid computing the inverse of a prohibitively large matrix, the ...

Certain Characteristics of Unlicensed Tree Expert Companies ...
(69.91%) found to be advertising tree services online were unlicensed tree experts (UnLTEs). .... other records in the listing of over 1,400 UnLTEs be- ... online directories of Verizon and AT&T telephone listings is ..... fónicas en Internet.

numerical approximation of the boundary control for the ...
Oct 11, 2006 - a uniform time T > 0 for the control of all solutions is required. This time T depends on the size of the domain and the velocity of propagation of waves. In general, any semi-discrete dynamics generates spurious high-frequency oscilla

Approximation of Dynamic, Stochastic, General ...
We also extend the method to solving Hansenks real business ...... r>. 1 #a '$4k>-$. % #' a$4n> % 4z>. 4w>. 1 4c>. 4g>. 1. )# !R δ$ !Κ ! G. 4 k>-$. % ξ ! W.

Algorithmic Computation and Approximation of ...
Some categories have multiple criteria to classify subcategories. ... measures that can be used to compute semantic similarity on other kinds of ontologies. ...... meaningful to use a σc threshold because in applications such as search engines,.

Legendre Pseudospectral Approximation of ...
are not known, at least not for large values of N. Instead, computer-codes have been developed ..... As can be seen, the ratios seem to support that the implemented method is of fourth-order accuracy in the time step ..... A, 365(2007), 278–283.

Capabilities of the Discrete Dipole Approximation for ...
solution of the system of 3Nd complex linear equations, where. Nd is the number of .... [9] M. A. Yurkin, V. P. Maltsev, and A. G. Hoekstra, “The discrete dipole.

Features and capabilities of the discrete dipole approximation code ...
4Computational Science research group, Faculty of Science, University of Amsterdam, ... M.Y. is supported by the program of the Russian Government “Research ... [2] M. A. Yurkin and A. G. Hoekstra, “The discrete dipole approximation: an ...