Error Analysis of Galerkin′s Method for Semilinear Equations
Abstract
We establish a general existence result for Galerkin′s approximate solutions of abstract semilinear equations and conduct an error analysis. Our results may be regarded as some extension of a precedent work (Schultz 1969). The derivation of our results is, however, different from the discussion in his paper and is essentially based on the convergence theorem of Newton’s method and some techniques for deriving it. Some of our results may be applicable for investigating the quality of numerical verification methods for solutions of ordinary and partial differential equations.
1. Introduction
Theorem 1.1 (see [1], Theorems 3.1 and 3.2.)One assumes (H1). Let R ∈ (0, ∞) be a constant and U = {u ∈ X; ∥u∥ ≤ R}. One assumes that φ : U → X is a completely continuous map such that φ(U) ⊂ U. Then the following holds.
- (i)
The equation has a solution uh in U∩Xh for any h and there exists a monotone decreasing sequence with limk→∞hk = 0 and u∞ ∈ U such that in X as k → ∞ and u∞ is a solution of (1.1). Moreover, if u∞ is the unique solution of (1.1) in U, then one has limh→0uh = u∞ in X.
- (ii)
Let u* ∈ U be a solution of (1.1). If φ has a Fréchet derivative, in a neighborhood, 𝒩, of u* and 0 is not in the spectrum of f′, then u* is the unique solution of (1.1) in 𝒩 and has a solution uh ∈ Xh for any h, which is unique for sufficiently small h and
(1.3)which means that ∥u* − uh∥ and ∥(I − Ph)u*∥ are equivalent infinitesimals as h → 0.
Various ordinary and partial differential equations appearing in mathematical physics can be written in the form (1.1) with (H2) under an appropriate setting of the functional spaces. See Section 5 for some concrete examples.
One of our motivations for this study is to investigate the quality of a numerical verification method for solutions of differential equations. Some of our results in this paper may be applicable for such a purpose. See Remark 2.7 for further information.
The paper is organized as follows. In Section 2 we describe our main results. We prepare some preliminary abstract results in Section 3 and apply them to prove our main results in Section 4. In Section 5 we present some concrete examples on semilinear elliptic partial differential equations.
Notations. Let 𝒳 and 𝒴 be Banach spaces.
- (1)
We denote by ∥·∥𝒳 the norm of 𝒳. If 𝒳 is a Hilbert space, then ∥·∥𝒳 stands for the norm induced by the inner product of 𝒳. For u ∈ 𝒳 and r ∈ (0, ∞), we write B𝒳(u ; r): = {v ∈ 𝒳; ∥v − u∥ < r}. The subscript will be often omitted if no possible confusion arises.
- (2)
For an open set V ⊂ 𝒳, C1(V, 𝒴) denotes the space of continuously differentiable functions from V to 𝒴.
- (3)
We denote by ℒ(𝒳, 𝒴) the space of bounded linear operators from 𝒳 to 𝒴 and ℒ(𝒳) stands for ℒ(𝒳, 𝒳). For T ∈ ℒ(𝒳, 𝒴), ∥T∥𝒳→𝒴 denotes the operator norm of T. The subscript will be omitted if no possible confusion arises.
- (4)
Let ϕ(h) and ψ(h) be nonnegative functions. We write ϕ(h) ~ ψ(h) if ϕ(h) and ψ(h) are infinitesimals of the same order as h → 0, that is, ϕ(h) = O(1)ψ(h) and ψ(h) = O(1)ϕ(h) as h → 0. We write ϕ(h)≃ψ(h) if ϕ(h) and ψ(h) are equivalent infinitesimals as h → 0, that is, ϕ(h) = {1 + o(1)}ψ(h) as h → 0.
- (5)
Let Ω be a bounded domain of Rn. We denote Lebesgue spaces by Lp(Ω)(1 ≤ p ≤ ∞) with the norms for. We denote by the completion of (the space of C∞ functions with compact support in Ω) in the Sobolev norm: . We denote by H−1(Ω) the Sobolev space with the norm . Here, 𝒟′(Ω) stands for the set of distributions on Ω.
2. Main Results
Proposition 2.1. There exist h* > 0 and such that the following (i)–(iii) hold.
- (i)
There exists R* > 0 such that u = uh is the only solution of fh(u) = 0 in B(u*; R*) for any h ∈ (0, h*).
- (ii)
u = uh is an isolated solution offh(u) = 0 for anyh ∈ (0, h*).
- (iii)
uh → u* in X as h → 0 with the estimate
(2.2)where and Ch → 1 as h → 0.
Remark 2.2. (i) Proposition 2.1(ii) is useful in our analysis below. Moreover, we immediately obtain from it that u = uh is an isolated solution of for any h ∈ (0, h*). This guarantees that we can always construct a Galerkin approximate solution uh by Newton’s method for small h > 0.
(ii) In various contexts in applications, Xh is finite-dimensional for any h. In such contexts the assumption (H1) implies that X is separable.
(iii) We do not assume dimXh < ∞. We briefly explain that it has some practical benefits. The case dimXh = ∞ appears, for example, in the following context. We are interested in the semi-discrete approximation to a periodic system described by a partial differential equation with a periodic forcing term. We may apply a Galerkin method only in space to the original system in order to construct a simpler approximate system described by ordinary differential equations. Then, for an isolated periodic solution of the original system, our Proposition 2.1 may guarantee that in a small neighborhood of it the approximate system has a periodic solution. For example, we can actually apply Proposition 2.1 to a semi-discrete approximation to a periodic system treated in [3]. See [4, Remark 3.4] for how to rewrite the system in [3] as (1.1).
In what follows in this section, always denotes the sequence as described in Proposition 2.1. Since u* − uh is decomposed into the Xh-component Phu* − uh and the -component (I − Ph)u*, we have and ∥(I − Ph)u*∥ ≤ ∥u* − uh∥. So, the last inequality and (2.2) immediately imply (2.3) below.
Corollary 2.3. We have
Actually, we easily verify that (2.3), (2.4) and (2.5) are mutually equivalent. They are very general features for the Galerkin method. The estimate (2.5) means that the Xh-component of the error ∥Phu* − uh∥ is an infinitesimal of a higher order of smallness with respect to the whole error ∥u* − uh∥ as h → 0.
The following two results are useful for applications (see Remark 2.7 below).
Theorem 2.4. We have the following:
Theorem 2.5. (i) We have
(ii) Let ɛh be a positive constant for h ∈ (0, h*) such that
In view of Theorem 2.5 (i) and (ii), we can always take in (2.10) such that εh → 0 as h → 0. The following Remarks 2.6 and 5.3 below shows that our estimate (2.10) is in general sharper than an estimate which can be derived directly from the discussion in [1].
Remark 2.6. (i) In the same way as in the proof of [1, Theorem 3.2] we can obtain an estimate related to (2.10). We set ηh∶ = (2ph + qh + rh)/(1 − (ph + qh + rh)), ph∶ = ∥A−1(I − Ph)T∥, qh∶ = ∥A−1PhT(I − Ph)∥ and rh∶ = ∥A−1∥ · ∥φ(uh) − φ(u*) − T(uh − u*)∥/∥uh − u*∥. It follows from Proposition 2.1 (iii) and Proposition 3.1 below that ph, qh and rh converge to 0 as h → 0. So, ηk → 0 as h → 0. Let be a positive constant for h ∈ (0, h*) such that . Then we have
(ii) When we compute for concrete examples (e.g., examples in Section 5 below), it seems reasonable to estimate qh as qh ≤ C∥T(I − Ph)∥. Here, C represents some positive constant independent of h. Then, it is actually necessary to take such that for small h > 0. On the other hand, roughly speaking, (2.9) means that we can take ɛh ≈ ∥T(I − Ph)∥ for small h > 0 (See Remark 5.3 below). (We note that Proposition 3.1 below implies that ∥T(I − Ph)∥ → 0 as h → 0.)
(iii) We consider the case where T is self-adjoint (e.g., Example 5.1 below). In this case, we have ∥(I − Ph)T∥ = ∥T(I − Ph)∥. So, by (2.12) is larger than for small h > 0.
Remark 2.7. We mention applications of our results. Some of our results may be applicable for testing the quality of a numerical verification algorithm for solutions of differential equations. In general we obtain an upper bound of ∥u* − uh∥ as output data from a numerical verification algorithm (See e.g., [5] and the references therein). By our Theorem 2.4 ∥u* − uh∥ is sufficiently close to ∥f(uh)∥ for sufficiently small h. So, Theorem 2.4 shows that we can check the accuracy of the output upper bound of ∥u* − uh∥ by finding the value of ∥f(uh)∥ when h is small. In [5] we proposed a numerical verification algorithm which also gives upper bounds of ∥Phu* − uh∥ as output data. Our Theorem 2.5 may be applicable for testing the accuracy of such upper bounds. See Remark 5.4 for more detailed information.
3. Preliminary Abstract Results
In this section, we prepare some abstract results in order to prove our main results in Section 2.
Proposition 3.1. We assume (H1). Let K : X → X be a compact operator. Then we have the following:
Proof. Though this result was proved in [6, Section 78], we give a simpler proof for the convenience of the reader. First we show that
Next, we describe some results in a more general setting. In what follows in this section, let 𝒳 and 𝒴 be Banach spaces and U ⊂ 𝒳 be an open set. We assume F ∈ C1(U, 𝒴).
Theorem 3.2. Let u0 ∈ U and L ∈ ℒ(𝒳, 𝒴) be bijective. We define a map g : U → 𝒳 by
Remark 3.3. (i) Theorem 3.2 is a new version of the convergence theorem of simplified Newton’s method, which is a refinement of the classical versions such as [5, Theorem 0.1]. Actually, the former implies the latter.
(ii) The convergence theorem of simplified Newton’s method is a very strong and general principle to verify the existence of isolated solutions. The reason is, roughly speaking, that the condition of the theorem is not only a sufficient condition to guarantee an isolated solution but also virtually a necessary condition for an isolated solution to exist. See [4, Remark 1.3] for the detail.
Proof of Theorem 3.2. Though we may consider Theorem 3.2 as a corollary of [5, Theorem 1.1], we describe the proof for completeness. We easily verify that u is a solution of F(u) = 0 if and only if u is a fixed point of g(u). Let u, v ∈ U. We obtain
The next result may be considered as a refinement of [7, Theorem 3.1 (3.14)] and [8, Theorem 3.1 (3.23)].
Proposition 3.4. Let u, v ∈ U , (1 − s)u + sv ∈ U for any s ∈ (0,1) and L ∈ ℒ(𝒳, 𝒴) be bijective. We set m : = maxs∈[0,1] ∥L − F′((1 − s)u + sv)∥. Then we have
Proof. The proof is similar to that of Theorem 3.2. Let g : U → 𝒳 be a map defined by (3.6). We have
Theorem 3.5. Let u = u* ∈ U be an isolated solution of the equation F(u) = 0. Let h0 > 0 be a positive constant, Fh ∈ C1(U, 𝒴) and Hh ∈ ℒ(𝒳, 𝒴)(0 < h < h0). We set H∶ = F′(u*). We assume that
- (a)
(3.22)
- (b)
u = uh is an isolated solution of Fh(u) = 0 for any h ∈ (0, h*),
- (c)
(3.23)
- (d)
Hh is bijective with ,
- (e)
the solution of Fh(u) = 0 is unique in B(u*; Rh) for any h ∈ (0, h*), where
(3.24) - (f)
(3.25)
Proof. By (3.20) and the stability property of linear operators (e.g., [3, Corollary 2.4.1]), and Hh are bijective for sufficiently small h > 0 and , in ℒ(𝒴, 𝒳) as h → 0. Let and . We set d(r): = d(r, u*) for r > 0 and define . Let ch : = 1/{1 − bh(2ηh)}, rh : = chηh and . Then, we easily verify that as h → 0,
4. Proofs of Main Theorems
We prove the results in Section 2. We use the notation (2.1).
Proof of Proposition 2.1. We apply Theorem 3.5 by putting 𝒳 = 𝒴 : = X, F : = f, Fh : = fh, H : = A and Hh : = Ah. We show (3.19)–(3.21). By (H1) we have fh(u*) = (I − Ph)u* → 0 in X as h → 0. Therefore, (3.19) holds. It follows from (H2) and Proposition 3.1 that
Proof of Theorem 2.4. We set u(s, h): = (1 − s)uh + su* for simplicity. Proposition 2.1 (iii) implies maxs∈[0,1]∥u* − u(s, h)∥ = ∥u* − uh∥→0 as h → 0. First we show (2.6). We have f(uh) = −(I − Ph)φ(uh) = (I − Ph)f(uh), in ℒ(X) as h → 0 and
Proof of Theorem 2.5. We set u*(s, h): = (1 − s)u* + sPhu* for simplicity.
- (i)
It follows from (H2) and Proposition 3.1 that
(4.6)
- (ii)
In the same way as (3.11) we have
(4.8)
Finally we derive (2.11) and (2.12).
Proof of (2.11) and (2.13). Without loss of generality we assume . First we derive (2.11). This proof is essentially the same as that of [1, Theorem 3.2]. It suffices to prove
5. Concrete Examples
We now present two examples.
Example 5.1. We consider the following Burgers equation:
Example 5.2. We consider the Emden equation
Remark 5.3. This remark is related to Remark 2.6.
(i) As mentioned in Remark 2.6 (ii), sups∈[0,1]∥φ′((1 − s)u* + sPhu*)(I − Ph)∥ ≈ ∥T(I − Ph)∥ holds in general. Actually, in Example 5.2 (resp., Example 5.1) our best possible upper bound of sups∈[0,1]∥φ′(u*(s, h))(I − Ph)∥ is the right-hand side of (5.13) (resp., (5.10)), which is just the same (resp., has the same order) as that of ∥T(I − Ph)∥.
(ii) We pointed out that our estimate (2.10) is in general sharper than (2.11), which is directly derived from the discussion in [1]. In order to show it concretely, we apply (2.11) to the equations in Examples 5.1 and 5.2. In both cases our best possible error estimate is the following:
Remark 5.4. Various numerical verification algorithms for solutions of differential equations were proposed up to now (see e.g., [10]). Some of them give upper bounds of ∥Phu* − uh∥ as output data (see [5]). Theorem 2.5 may be applicable for checking the accuracy of such output upper bounds since we can apply it to given problems in order to compute the concrete order of ∥Phu* − uh∥ as h → 0. For example, we treated problems (5.5) and (5.11) as concrete numerical examples in [5], where we proposed a numerical verification algorithm based on a convergence theorem of Newton’s method. In these problems (5.3) is the theoretical estimate of ∥Phu* − uh∥ derived from our Theorem 2.5. The output data as upper bounds of ∥Phu* − uh∥ in [5, Section 3] seem to have just the order of h2 as h → 0. So, the accuracy of such output upper bounds in [5, Section 3] is satisfactory as long as we judge it by the theoretical estimate (5.3).
Acknowledgments
The author would like to express his sincere gratitude to Professor Takuya Tsuchiya and Professor Atsushi Yagi for their valuable comments and encouragement. He is grateful to the referee for constructive comments.