• Keine Ergebnisse gefunden

Scalings and singular limits on the basis of the D1P2 model

2.2 A singularly perturbed initial value problem

This observation shows, that we cannot obtain from (2.24) a unique φ and thus unique population functions f1,f2, without having some extra information about their initial values.

Remark 2.3. (Periodic BC)The compatibility condition (2.25) emerges also in the case ofperiodicboundary conditionsf(t,0) =f(t, L), implying immediately peri-odic boundary conditions for both11moments: u(t,0) =u(t, L) andφ(t,0) =φ(t, L).

In order to get classical periodic solutions, the derivatives occurring in the evolution equations are also required to be continuous periodic functions.

As far as the construction of a periodic φ is concerned (proof of proposition 2.1), we recognize in the explicit formula (2.27), that the only term12 making trouble is the first integral withxas upper integration bound. In fact, the periodicity, i.e. the equality for x = 0 and x = L enforces the integral to vanish over the full space domain [0, L], leading thus to the compatibility condition. The constant C can be chosen arbitrarily.

2.2 A singularly perturbed initial value problem

In the previous section, two scalings for lattice-Boltzmann equations were presented using the D1P2 velocity model as example. Heuristically, it was shown how the scalings give rise to different target equations. In this section we want to inspect the limit more carefully, which is generated by the parabolic scaling. To start with, we take up the the scalar equation (2.21) to avoid the lattice-Boltzmann system.

Further simplifications are obtained by considering the case of a trivial flux function f and absent sources q. Under these restrictions the IBVP (2.24) over the spatial domainX = [0, L] with homogeneous Dirichlet boundary conditions reads

EQ : ǫ2τ ∂2tuǫ+∂tuǫ−τ ∂x2uǫ = 0, τ >0 BC : uǫ(·,0) = 0 =uǫ(·, L)

IC : uǫ(0,·) =g ∧ ∂tuǫ(0,·) =h





. (2.29)

Here we want to focus on the behavior of the solution uǫ, if the parameter ǫtends to zero. This is interesting, since the equation is losing its highest time derivative for ǫ = 0. Thereby the equation undergoes an essential qualitative change13: the second initial condition prescribing∂tuǫ(0,·) gets redundant (resulting in an over-determination by too many initial conditions) due to an abrupt alteration of the equation type (hyperbolic→ parabolic). Such a limit is called singular.

11Note the contrast to the bounce-back type boundary conditions affecting only one moment.

12Observe that the second integral is with respect to the time. That is why the assumed peri-odicity ofuandxucan directly pass over toφ.

13For example: second order differential equations are characterized by an oscillating behavior of their solutions (e.g. wave equation, harmonic oscillator), whereas solutions to first order equations exhibit rather a monotonic drop or growth (e.g. heat equation, radioactive decay). Furthermore, time evolution operators of parabolic equations are characterized by smoothing the regularity of the initial condition; this feature is not shared by hyperbolic equations. We will be confronted with this property in the course of this section.

In view of these malignancies it is by far not self-evident that the solution uǫ con-verges to the solution of the natural limit system,

EQ : ∂tu−τ ∂x2u= 0 BC : u(·,0) = 0 =u(·, L) IC : u(0,·) =g



 (2.30)

wheregis the same as in (2.29). This is all the more surprising, as the second initial condition does not need to be “consistent” with the limit, i.e. convergence happens even ifh6=∂tu(0,·). In the sequel we refer to (2.29) as theperturbed problem, while (2.30) denoted as the limit problem.

In this section we are going to prove the convergence of uǫ→u employingFourier series, that permit a quasi analytical computation of the solution for both IBVPs.

Alternatively it is possible to approximateuǫ by a truncated expansion of the form uǫ = u(0) +ǫu(1) +...+ǫku(k)+ O(ǫk+1), finally using a stability result to infer convergence from consistency. Since this approach (which applies undoubtedly to a wider range of problems) is pursued in the subsequent chapters, we prefer here the classical Fourier technique for comparison14.

2.2.1 The Fourier coefficient functions Solution of the limit problem

IBVPs forlinearPDEs are frequently attacked by an ansatz that is calledseparation of variables. The idea consists in writing the solution as a product, where one fac-tor depends only on time while the other is purely space-dependent. This approach decomposes the original problem into a boundary eigenvalue-problem (BEVP) for the spatial function and into an initial value problem for the time-dependent part.

Of course, not every solution of the original IBVP is so simply structured that it factorizes with respect to time and space. But thanks to the linearityof the prob-lem, simple multiplicative solutions can be linearly combined, to synthesize more complicated solutions. In fact, under appropriate conditions, any solution can be expanded by means of simple solutions. Thus we arrive directly at what is known under the name of generalized Fourier series.

The solution of the IBVP (2.30) leads to the BEVP of the one-dimensional Laplacian (second derivative) with homogeneous Dirichlet boundary conditions. The eigen-functions sn to this problem are given by the sine

sn(x) :=

q2

Lsin(nπx/L) for n∈N (2.31)

having an integral number of bellies between the boundary points 0 and L. More precisely these functions satisfy the equation

−τ ∂x2snnsn with the eigenvalues λn:=τ L2

. (2.32)

14Apparently, the Fourier approach can be done with weaker regularity requirements.

2.2. A singularly perturbed initial value problem 83

The eigenfunctions sn are normalized and pairwise orthogonal. Furthermore the set of eigenfunctions is complete in the Hilbert spaceL2(0, L).

Let us suppose that the initial conditiongis inL2(0, L). Then it can be represented by a Fourier series

g=X

nN

αnsn with αn:=

Z L 0

g(x)sn(x) dx . (2.33) The coefficientsαn are denoted as theinitial Fourier coefficients.

Theorem 2.1. Forg∈ L2(0, L) the IBVP (2.30) has a unique solution u∈ C [0,∞),L2(0, L)

∩ C1,2 (0,∞)×[0, L]

given by u(t, x) =X

nN

σn(t)sn(x) with σn(t) :=αneλnt. (2.34) Moreover u∈ C (0,∞)×[0, L]

.

The proof of this theorem can be performed by means of Fourier series where each coefficientσn comes out as solution of the IVP

EQ : σ˙nnσn= 0 IC : σn(0) =αn

)

. (2.35)

Fourier coefficient functions for the perturbed problem

Since the spatial differential operator occurs in the perturbed equation (2.29) as well as in the limit equation (2.30), we will work with its eigenfunctions (2.31) to solve (2.29) too.

Let us assume that the second initial condition in (2.29) admits also a Fourier expansion

h=X

nN

βnsn with βn:=

Z L 0

h(x)sn(x) dx . (2.36) In analogy to (2.34) we start with the ansatz

uǫ(t) =X

nN

σǫ,n(t)sn(x). (2.37)

Plugging this expansion into (2.29), a formal computation yields the following IVP for each coefficient:

EQ : ǫ2τσ¨ǫ,n+ ˙σǫ,nnσǫ,n= 0 IC : σǫ,n(0) =αn ∧ σ˙ǫ,n(0) =βn

)

(2.38) In order to set up the fundamental system we need the zeros of the associated characteristic polynomial

z7→ǫ2τ

z22ωz+ǫ2ωλn

. (2.39)

These are given by ψǫ,n := −12ǫ2ω+

q1

4ǫ4ω2−ǫ2ωλn = −12ωǫ2h 1−p

1−4ǫ2τ λni

, (2.40) χǫ,n := −12ǫ2ω−

q1

4ǫ4ω2−ǫ2ωλn = −12ωǫ2h 1 +p

1−4ǫ2τ λni

. (2.41) It will be convenient to abbreviate the square root:

Wǫ,n:=p

1−4ǫ2τ λn

( ψǫ,n=−12ωǫ2[1−Wǫ,n]

χǫ,n=−12ωǫ2[1 +Wǫ,n] . (2.42) In general, two cases have to be distinguished concerning the fundamental system:

1) two zeros ψǫ,n6=χǫ,n⇔Wǫ,n6= 0 : t7→eψǫ,nt, t7→eχǫ,nt 2) double zero ψǫ,nǫ,n⇔Wǫ,n= 0 : t7→eψǫ,nt, t7→teψǫ,nt

The first case is the standard situation. If both zeros coincide, a degenerated case is obtained.

In order to avoid annoying repetitions of case distinctions, we henceforth adopt the subsequent agreement till explicit revocation.

Assumption 2.1. The real parameter ǫ 6= 0 is thought as a null-sequence (con-verging towards zero) subject to the sole condition15that it has no common element with the null-sequence n7→ 2τ πnL , i.e.:

∀n∈N: ǫ6= 2L

τ πn ⇔ 1−4ǫ2τ λn6= 0.

This condition implies Wǫ,n 6= 0, such that we are always in the regular case with two different zeros of the characteristic polynomial. Hence, the Fourier coefficient functionσǫ,n is a linear combination of eψǫ,nt and eχǫ,ntrespecting the initial condi-tions in (2.38). A short calculation yields the formula

σǫ,n(t) := βn−αnχǫ,n

ψǫ,n−χǫ,n eψǫ,nt + αnψǫ,n−βn

ψǫ,n−χǫ,n eχǫ,nt. (2.43) Although we have derived this expression for σǫ,n starting with the ansatz (2.37), a rigorous treatment requires that we take this equation as a definition16.

Sorting for terms containing either αn orβn we get σǫ,n(t) =αnψǫ,neχǫ,nt−χǫ,neψǫ,nt

ψǫ,n−χǫ,n + βneψǫ,nt−eχǫ,nt

ψǫ,n−χǫ,n . (2.44)

15Notice thatǫoccurs always in squared form, that is why it is redundant to requireǫ >0.

16We do not yet know whether the resulting Fourier series will be convergent inL2(0, L), not to mention any regularity properties which are necessary to justify the formal computation. Therefore we are obliged to go the other way round. So it has to be shown that the coefficients defined by the IVP (2.38) give rise to a converging Fourier series with a sufficiently smooth limit function.

2.2. A singularly perturbed initial value problem 85

Furthermore using the relation ψǫ,n−χǫ,n=ωWǫ,n2, this is transformed into σǫ,n(t) = 12αnh

(1−Wǫ,n1)eχǫ,nt + (1 +Wǫ,n1)eψǫ,nti

+ 12βnǫ2τ Wǫ,n1

eψǫ,nt−eχǫ,nt

= 12αnh

eψǫ,nt+ eχǫ,nt

+Wǫ,n1 eψǫ,nt−eχǫ,nti

+12βnǫ2τ Wǫ,n1

eψǫ,nt−eχǫ,nt

nρn,ǫn̺n,ǫ , (2.45)

where we have set

ρǫ,n(t) := 12

eψǫ,nt+ eχǫ,nt

+ 12Wǫ,n1

eψǫ,nt−eχǫ,nt], (2.46)

̺ǫ,n(t) := 12ǫ2τ Wǫ,n1

eψǫ,nt−eχǫ,nt

. (2.47)

In order to prove boundedness of these coefficient functions we prefer another rep-resentation. For this, recall the definition of cosh(z) = 12(ez+ ez) and sinh(z) =

1

2(ez−ez). Using the representation ofψǫ,n, χǫ,n in (2.42) we get:

eψǫ,nt+ eχǫ,nt = e12ωt/ǫ2+12ωWǫ,nt/ǫ2 + e12ωt/ǫ212ωWǫ,nt/ǫ2

= e12ωt/ǫ2h

e12ωWǫ,nt/ǫ2 + e12ωWǫ,nt/ǫ2i

= 2e12ωt/ǫ2cosh 12ωWǫ,nt/ǫ2 . Similarly we obtain:

eψǫ,nt−eχǫ,nt = e12ωt/ǫ2h

e12ωWǫ,nt/ǫ2 −e12ωWǫ,nt/ǫ2i

= 2e12ωt/ǫ2sinh 12ωWǫ,nt/ǫ2 .

This finally gives the following representation for the coefficient functions defined in (2.46) and (2.47)

ρǫ,n(t) = e12ωt/ǫ2cosh 12ωWǫ,nt/ǫ2

+ e12ωt/ǫ2Wǫ,n1sinh 12ωWǫ,nt/ǫ2

, (2.48)

̺ǫ,n(t) =ǫ2τe12ωt/ǫ2Wǫ,n1sinh 12ωWǫ,nt/ǫ2

. (2.49)

Boundedness of the Fourier coefficient functions

The boundedness of the Fourier coefficient functions may be related to stability estimates, that are necessary for alternative approaches (e.g. regular expansions).

The subsequent lemma represents a preparative and indispensable intermediate step for proving further results.

Lemma 2.1. The coefficient functions ρǫ,n defined in (2.46) satisfy the uniform estimate

ǫ,n(t)| ≤2 for allt≥0, n∈N and all admissible ǫ.

Proof: From (2.48) we obtain:

ǫ,n(t)| ≤

e12ωt/ǫ2cosh 12ωWǫ,nt/ǫ2

| {z }

=:Aǫ,n(t)0

+

e12ωt/ǫ2Wǫ,n1sinh 12ωWǫ,nt/ǫ2

| {z }

=:Bǫ,n(t)0

.

We show in the sequel that Aǫ,n(t), Bǫ,n(t) ≤ 1 independently of ǫ, n and t, such that |ρǫ,n(t)| ≤Aǫ,n(t) +Bǫ,n(t) = 2. There are two cases to be distinguished:

I) 1−4ǫ2τ λn>0⇔Wn,ǫ∈R due to λn, τ, ǫ >0 follows: 0< Wǫ,n<1 II) 1−4ǫ2τ λn<0⇔Wn,ǫ∈iR there exists wǫ,n∈R+ with: Wǫ,n= iwǫ,n Here ‘i’ denotes theimaginary unit.

Ad I) 0< Wǫ,n<1

Aǫ,n(t) =

e12ωt/ǫ2cosh 12ωWǫ,nt/ǫ2

≤ e12ωt/ǫ2cosh 12ωt/ǫ2

= 12e12ωt/ǫ2

e12ωt/ǫ2+ e12ωt/ǫ2

= 12 +12eωt/ǫ2

12 +12 = 1

For the first inequality we have employed thatWǫ,nis positive and that cosh is monotonically increasing for positive arguments. The second inequality holds fort≥0 and ω >0.

In order to estimate Bǫ,n(t) (note remark 2.4) we consider the function f(a, x) := easinh(ax)

x for 0≤x≤1 ∧ a≥0.

We assert that for a fixeda >0 the function (0,1)∋x7→f(a, x) is monotoni-cally increasing. As the function is continuously differentiable, this is verified by checking the non-negativity of its derivative. Indeed:

xf(a, x) = eaxacosh(ax)−sinh(ax)

x2 >0 (2.50)

Due to its monotonicity the function (0,1)∋x7→f(a, x) attains its extrema at the boundary i.e. for xtending to 0 or 1.

limx0f(a, x) = ealim

x0

sinh(ax) x

L’Hospital

= ealim

x0

acosh(ax)

1 = eaa≤e1 < 12 limx1f(a, x) = easinh(a) = 12ea ea−ea

= 12(1−e2a)≤ 12

Observe that the function a 7→ aea is maximal for a = 1 (see lemma 2.8). Since a ≥ 0 is arbitrary (although fixed) we get |f(a, x)| ≤ 12 < 1 for 0< x <1. Identifyingx withWǫ,nand setting a= 12ωt/ǫ2 we obtain

Bǫ,n(t) =

e12ωt/ǫ2Wǫ,n1sinh 12ωWǫ,nt/ǫ2 <1.

2.2. A singularly perturbed initial value problem 87

The correctness of equation (2.50) ensues from the fact thatycosh(y)>sinh(y) for ally >0. As the hyperbolic functions are real analytic, this can easily be seen by writing down the Taylor series:

ycosh(y) = y 1 +2!1y2+4!1y4+6!1y6+...

=y+2!1y3+4!1y5+6!1y7+...

sinh(y) = y+3!1y3+5!1y5+7!1y7+... . Ad II) Wǫ,n= iwǫ,n, wǫ,n>0

In the following estimates we exploit the identities sinh(ix) = i sin(x) and cosh(ix) = cos(x) for real x.

Aǫ,n(t) =

e12ωt/ǫ2 ·cosh 12ωWǫ,nt/ǫ2

= e12ωt/ǫ2·

cosh 12ωiwǫ,nt/ǫ2

= e12ωt/ǫ2·

cos 12ωwǫ,nt/ǫ2

≤ e12ωt/ǫ2·1

≤ 1

The first inequality results from the fact, that the cosine is bounded for real arguments in contrast to the cosine hyperbolicus. Therefore the exponential function, that was important for case I to damp the cosine hyperbolicus, does not play a crucial role this time.

Bǫ,n(t) =

e12ωt/ǫ2·Wǫ,n1·sinh 12ωWǫ,nt/ǫ2

= e12ωt/ǫ2 · |iwǫ,n|1·

sinh 12ωiwǫ,nt/ǫ2

= e12ωt/ǫ2 ·wǫ,n1·

i·sin 12ωwǫ,nt/ǫ2

= e12ωt/ǫ2 ·

˛

˛

˛sin1

2ωwǫ,nt/ǫ2”˛

˛

˛ wǫ,n

≤ e12ωt/ǫ2 ·12ωt/ǫ2

≤ 1

The first inequality emanates from the mean value theorem, which gives sin 12ωwǫ,nt/ǫ2

= 12ωt/ǫ2 ·cos 12ωξǫ,nt/ǫ2

·wǫ,n for some 0 < ξǫ,n < wǫ,n. As the cosine function and the mapping a 7→ aea are bounded by 1 for non-negative arguments, we obtain the final estimate.

Remark 2.4. Under condition I) the mean value theorem is of no use to estimate Bǫ,n, since we get here a cosine hyperbolicus instead of a cosine. Notice, that the occurring exponential exp(−12ωt/ǫ2) can only be utilized to “slay” either the cosine hyperbolicus or the factor 12ωt/ǫ2, which would come into play by the mean value theorem.

Corollary 2.1. The coefficient functions̺ǫ,ndefined in (2.47) satisfy the uniform estimate

ǫ,n(t)|< τ ǫ2 for all t≥0, n∈Nand all admissible ǫ.

Proof: The representation of ̺ǫ,n in (2.49) corresponds to the second addend of (2.48) multiplied additionally by ǫ2τ. Therefore̺ǫ,n is estimated in the same way

as Bǫ,nin the proof of the previous lemma.

Summarizing, we have obtained

ǫ,n(t)|<2|αn|+|βn|τ ǫ2 . (2.51) The uniform boundedness of the coefficient functions ρǫ,n and ̺ǫ,n has two useful consequences.

Lemma 2.2. The Fourier coefficient functions σǫ,n(t) represent square-summable sequences with respect to n. Furthermore theirℓ2(N)-norm can be bounded indepen-dently17 of 0< ǫ≤1 and t≥0.

Proof: Recall that the sequences (αn)nN and (βn)nN are square-summable as Fourier coefficients of the L2-functions g, h. Introducing A2 := P

n1n|2 and B2 := P

n1n|2 and exploiting the estimates in lemma 2.1 and corollary 2.1 we obtain:

X

n1

ǫ,n(t)|2 (2=.45)X

n1

αnρǫ,n(t) +βn̺ǫ,n(t) 2

≤X

n1

n|2ǫ,n(t)|2+ 2X

n1

n||βn||ρǫ,n(t)||̺ǫ,n(t)|+X

n1

n|2ǫ,n(t)|2

≤4X

n1

n|2+ 4τ ǫ2X

n1

n||βn|+τ2ǫ4X

n1

n|2

≤4A2+ 4τ ǫ2AB+τ2ǫ4B2 <4A2+ 4τ AB+τ2B2<∞ for ǫ≤1 Observe that we have employed theCauchy-Schwarz inequality for infinite sums.to estimate the product sum P

n1n||βn|.

The sequences composed of the time-dependent Fourier coefficient functions permit not only a common bound for their ℓ2-norms (independent of ǫ), but their sums converge alsouniformly. The statement of the next lemma makes this more precise.

Lemma 2.3. For every δ >0 there exists N =N(δ)∈N, such that X

n>N

ǫ,n(t)|2 < δ independently of t≥0 and 0< ǫ≤1.

17The conditionǫ1 in the assumptions of this and the next lemma seems to be artificial and arbitrary, since other upper bounds are possible as well. Although the condition might appear superfluous, as we consider the limitǫ0, we have added it here to be formally correct.

2.2. A singularly perturbed initial value problem 89

Proof: We remember again that (αn)nN and (βn)nN are square-summable. On account of the Cauchy-Schwarz inequality, the sequence (αnβn)nN is summable.

Hence for every η >0 there exist natural numbers Nα, Nβ, Nαβ ∈Ndepending on estimate in the proof of the preceding lemma:

X

Remark 2.5. This lemma turns out to be the crucial tool in dealing with the infinite Fourier series for uǫ, that we have formally written down in (2.37). It permits us to show that the Fourier coefficients functions σǫ,n bequeath properties – like continuity in time or convergence with respect toǫ– to the whole series.

Estimate of the time derivative

In subsection 2.2.3 we need an estimate similar to (2.51) also for ˙σǫ,n, that shall be anticipated here.

Lemma 2.4. The time derivative of the Fourier coefficient functions satisfies the estimate

|σ˙ǫ,n(t)| ≤ |αnn+ 2|βn| for all t≥0, n∈N and any admissible ǫ >0.

Proof: Computing the time derivative of the definition (2.43)

˙

we recognize in the first addend the coefficient function ̺ǫ,n (see (2.47)), while the second addend contains a factor ˜ρǫ,n being similar to the coefficient function ρǫ,n (see (2.46)). With

we obtain

˙

σǫ,n(t) =αnωǫ2λn̺ǫ,n(t) +βnρ˜ǫ,n(t).

Applying corollary (2.1), we find the estimate

|σ˙ǫ,n(t)| ≤ |αn|ωǫ2λnǫ,n(t)| + |βn||ρ˜ǫ,n(t)|

≤ |αnn+|βn||ρ˜ǫ,n(t)|. (2.52) In order to estimate |ρ˜ǫ,n(t)|, a further computation using ψǫ,n−χǫ,n=ωWǫ,n2 yields:

˜

ρǫ,n(t) = 12(1−Wǫ,n1)eψǫ,nt+ 12(1 +Wǫ,n1)eχǫ,nt

= 12 eψǫ,nt−eχǫ,n

+ 12Wǫ,n1 eχǫ,nt+ eψǫ,n

= e12ωt/ǫ2cosh(12ωWǫ,nt/ǫ2) − Wǫ,ne12ωt/ǫ2sinh(12ωWǫ,nt/ǫ2). A comparison with (2.46) shows, that ˜ρǫ,n is the difference of those terms, which give as sum ρǫ,n. Since each of the two terms has been estimated separately in lemma 2.1, we conclude

|ρ˜ǫ,n(t)| ≤2.

With (2.52) this gives the assertion.

Convergence of the Fourier coefficient functions for ǫց 0

Let us first consider how the zeros in (2.40) and (2.41) behave, whenǫtends towards zero. The characteristic polynomial is of the form

p(z) =ǫ2z2+az+b witha, b∈R\ {0}. (2.53) Obviously the quadratic polynomial degenerates to a linear function for vanishing ǫ. Its zeros are given by

z+ǫ :=

1 2a+

q1 4a2ǫ2b

ǫ2 and zǫ:=

1 2a

q1 4a2ǫ2b

ǫ2 . (2.54)

Observe that both zeros are real for sufficiently small ǫ.

Lemma 2.5. The zeros in 2.54 have the following asymptotic comportment:

i) a >0 : zǫ+−−→ −ǫ0 ba and zǫ−−→ −∞ǫ0 ii) a <0 : zǫ+−−→ ∞ǫ0 , and zǫ−−→ −ǫ0 ab

Proof: ad i) As the square root is by definition positive (for sufficiently small ǫ) and asa >0 by hypothesis, we see that the numerator ofzǫconverges to−a, while the positive denominator ǫ2 tends to 0. Hencezǫ goes to minus infinity.

In order to compute the limit of z+ǫ we use L’Hospital’s rulewith f(ǫ) :=−12a+

q1

4a2−ǫ2b and g(ǫ) :=ǫ2.

2.2. A singularly perturbed initial value problem 91

Since f(0) = 0 andg(0) = 0 we compute the first derivatives:

f(ǫ) =−ǫb 14a2−ǫ2b12

and g(ǫ) = 2ǫ

Once more we obtain f(0) = 0 and g(0) = 0, such that we have to pass over to the second derivatives:

f′′(ǫ) =−b 14a2−ǫ2b12

−ǫ2b(14a2−ǫ2b)32 and g′′(ǫ) = 2 Now we get: lim

ǫ0z+ǫ = lim

ǫ0 f(ǫ) g(ǫ) = lim

ǫ0 f(ǫ) g(ǫ) = lim

ǫ0 f′′(ǫ)

g′′(ǫ) = fg′′′′(0)(0) = b(a2/4)

1 2

2 =−ba. ad ii) The second case comes out in complete analogy where the term (a2/4)12 must be evaluated with care, as it is −a/2 for a <0.

The finite limit −ab is just the zero of the affine-linear function, that is obtained by setting ǫ= 0 in (2.53).

−30 −25 −20 −15 −10 −5 0

−4

−3

−2

−1 0 1 2 3 4

Re

Im

Figure 2.1: The figure illustrates the position of the characteristic rootsψǫ,n(red dots) andχǫ,n (blue crosses) in the complex plane for different values ofǫ. For this example the parameters areL= 1,τ = 0.1, n= 2whileǫ varies between12 and 0.55 in steps of0.005. For relatively largeǫthe roots are complex. Due to the real coefficients of the characteristic polynomial (2.39), ψǫ,n andχǫ,n are complex conjugated to each other.

If ǫ falls below a certain threshold value, ψn,ǫ converges on the real axis from left to

−λn (thick bullet) while χǫ,ntends monotonically to −∞.

Let us apply the lemma to the case of our interest. By identifying awith ω and b withωλnwe conclude immediately:

Corollary 2.2. The roots (2.40) and (2.41) of the characteristic polynomial asso-ciated to (2.38) display the following asymptotic behavior:

ψǫ,n−−→ −ǫ0 λn and χǫ,n−−→ −∞ǫ0 .

Now we are enabled to verify the convergence of the Fourier coefficient functions σǫ,n. Besides the boundedness and the presented implications, this is the second ingredient entering the convergence proof of theorem 2.3.

Lemma 2.6. If ǫ tends to zero, the Fourier coefficient functions σǫ,n defined in (2.43) converge pointwise for all t ≥ 0 to the Fourier coefficient functions σn(t) pertaining to the solution of the limit problem (2.30):

σǫ,n(t)−−→ǫ0 σn(t) =αneλnt.

Proof: For t = 0 the initial condition yields: σǫ,n(0) = αn = σn(0). So the case t > 0 remains to be considered. Let n∈Nbe fixed. Since χǫ,n converges to minus infinity for ǫ↓ 0, there is ǫ00(n) such that χǫ,n 6= 0 for all admissible ǫ < ǫ0. As all limits exist, we are allowed to apply thelimit rulesfor arithmetic operations:

σn(t) = lim

Multiplying denominator and numerator by χǫ,n yields (2.43) and hence the

asser-tion.

2.2.2 Solution of the perturbed problem and convergence

In (2.37) we have chosen for uǫ a Fourier series ansatz without knowing, whether this will lead to a reasonable result. In order to impart a meaning to (2.37) and to justify thereby the ansatz, we have to show that the sequence is convergent in some sense. This is the intention of the next proposition.

Proposition 2.2. Let the coefficient functions σǫ,nbe as in (2.43). Then uǫ(t, x) :=X

nN

σǫ,n(t)sn(x)

defines for all t≥0 a square-integrable function, i.e.uǫ(t,·)∈ L2(0, L).

Proof: Generally, a series of pairwise orthogonal elements converges in a Hilbert space if and only if the sum of the their squared norms is convergent. As thesn’s are normalized in L2(0, L), we have to check, whether the sum of the squared moduli

ǫ,n(t)|2 is finite. But this is just affirmed by lemma 2.2.

2.2. A singularly perturbed initial value problem 93

The proposition does not yet classify uǫ as a function of time. This is the object of the next theorem.

Theorem 2.2. The functionuǫ defined in proposition 2.2 depends continuously on the time, more exactly:

uǫ∈ C [0,∞),L2(0, L)

As already insinuated in remark 2.5, the proof of this theorem relies on lemma 2.3.

Since we do not dispose of equicontinuity in time for the whole sequence (σn,ǫ)nN, the Fourier series representing the differenceuǫ(t,·)−uǫ(t0,·) is split into two parts.

The first portion comprises only finitely many addends up to a certain index. These terms are handled by exploiting the continuity of the coefficient functions to be transferred on the whole series. The tail of the series, is controlled in a general way by means of the uniform decay provided by lemma 2.3.

Proof: We have to show that for all t0 ∈ [0,∞) and for any δ > 0 there exists a real numberθ >0, such that

kuǫ(t,·)−uǫ(t0,·)k22 < δ if |t−t0|< θ.

UsingParseval’s theorem and the orthonormality of thesn’s , the squaredL2-norm can be expressed in terms of the Fourier coefficient functions:

kuǫ(t,·)−u(t,·)k22 = X

n1

σǫ,n(t)−σǫ,n(t0)

·sn 2

2

=X

n1

σǫ,n(t)−σǫ,n(t0)

2· ksnk22

=X

n1

σǫ,n(t)−σǫ,n(t0)

2 . (2.55)

Lemma 2.3 guarantees the existence of M =M(δ)∈N, such that X

n>M

ǫ,n(t)|2 < δ3 and X

n>M

ǫ,n(t0)|2 < δ3 .

As each of the coefficient functions is continuous in time, we can find real numbers θ1, ..., θM >0 with

ǫ,n(t)−σǫ,n(t0)|< 3Mδ if |t−t0|< θn and n∈ {1, ..., M}.

Choosing finallyθ:= min{θ1, ..., θM}we obtain from (2.55) by applying the triangle inequality

kuǫ(t,·)−u(t,·)k22 ≤ XM n=1

σǫ,n(t)−σǫ,n(t0)

2+ X

n>M

ǫ,n(t)|2+ X

n>M

ǫ,n(t0)|2

≤M 3Mδ +δ3+ δ3

whenever |t−t0|< θ. Note that θ1, ..., θM and hence θ dependa priori (for given initial Fourier coefficients defining σǫ,n) on the choice of δ and M. If one tries to choose M as small as possible, M depends on δ and thereforeθ depends actually

only on δ.

At this point it would be natural to expect a statement establishinguǫ as the solu-tion of the perturbed problem (2.29). However, in opposisolu-tion to the limit problem (2.30), there is no result equivalent to theorem 2.1. The reason18 for the different behavior lies in the second time derivative that makes (2.29, EQ) to a sort of wave equation, that is of hyperbolic nature despite the attenuating action of the first time derivative. It is a characteristic property of hyperbolic equations that their associated time evolution operator does not smooth the regularity of the initial data. In contrast, the solution of a parabolic equation like the diffusion equation in (2.30) acquires C-regularity as soon as t >0 – even if started by an L2-function.

Therefore it is not possible to assign touǫ the meaning of a classical solution for the perturbed IVBP under the assumption of L2 initial data. Nevertheless it is possi-ble to interpret uǫ as solution in a wider, generalized sense (weak or distributional solution).

In view of these facts, the convergence of uǫ towards the C-solution u of the limit problem might appear questionable. On the other hand, having verified the

In view of these facts, the convergence of uǫ towards the C-solution u of the limit problem might appear questionable. On the other hand, having verified the