• No results found

The number of boundary conditions for initial boundary value problems

N/A
N/A
Protected

Academic year: 2021

Share "The number of boundary conditions for initial boundary value problems"

Copied!
10
0
0

Loading.... (view fulltext now)

Full text

(1)

The Number of Boundary

Conditions for Initial

Boundary Value Problems

Jan Nordström and Thomas Hagstrom

(2)

The Number of Boundary Conditions for

Initial Boundary Value Problems

J A N N O R D S T R ¨

O M

1

A N D

T H O M A S H A G S T R O M

2

1

Department of Mathematics, Link¨oping University, SE-581 83 Link¨oping, Sweden

2Department of Mathematics, Southern Methodist University, Dallas, TX 75275 USA

(Received 8 February 2020)

Both the energy method and the Laplace transform method are frequently used for determining the number of boundary conditions required for a well posed initial boundary value problem. We show that these two distinctly different methods yield the same results. Key words: initial boundary value problems; number of boundary conditions; energy method; Laplace transform method; normal mode analysis; Navier-Stokes equations

1. Introduction

It is common to employ either the energy method or the Laplace transform method in order to determine how many boundary conditions an initial boundary value problem require for well-posedness. The energy method builds on integration-by-parts (see Kreiss & Lorenz (1989), Gustafsson et al. (1995), Gustafsson & Sundstr¨om (1978), Nordstr¨om & Sv¨ard (2005), Nordstr¨om (2017)). One obtains an expression for the energy rate that involves boundary terms. The number of boundary conditions is given by the minimal number that limits these boundary terms.

The Laplace transform method is different, and employs an expansion of the solution in modes (see Hersh (1963), Hersh (1964), Kreiss (1970), Strikwerda (1977), Sakamoto (1982), Engquist & Gustafsson (1987), Eriksson & Nordstr¨om (2017)) that turn the initial boundary value problem into a one-dimensional boundary value problem. The number of boundary conditions is given by the number of conditions required to determine these modes. These two methods, give the same result on well-known equation sets (see Nordstr¨om (1989), Nordstr¨om & Gustafsson (2003), Nordstr¨om et al. (2007), Lauren & Nordstr¨om (2018)). In this paper we explain why.

2. Preliminaries

We start by presenting the general system of equations that will be investigated, and provide basic results for the Cauchy problem.

2.1. An incompletely parabolic system of equations

The three-dimensional constant coefficient system of equations that we will consider is Vt+ ¯AVx+ ¯BVy+ ¯CVz= ( ¯D11Vx+ ¯D12Vy+ ¯D13Vz)x+ ( ¯D21Vx+ ¯D22Vy+ ¯D23Vz)y

+ ( ¯D31Vx+ ¯D32Vy+ ¯D33Vz)z. (2.1)

(3)

2 J. Nordstr¨om and Thomas Hagstrom

In (2.1), V is the solution, ¯A, ¯B, ¯C are symmetric matrices related to the hyperbolic terms and the parabolic matrices ¯D11, ¯D22, ¯D33, ¯D12+ ¯D21, ¯D13+ ¯D31, ¯D23+ ¯D32 are

symmetric. Furthermore, the structure of the matrices ¯Dij in block form is

¯ Dij =  (Dij)mxm 0mxn 0nxm 0nxn  , (2.2)

where subscripts indicate the size of the matrices. Under the assumptions on the matrices Dij listed in (2.5) below this yields an incompletely parabolic system.

Remark 1. Abarbanel & Gottlieb (1981) showed that the three-dimensional compress-ible Navier-Stokes equations with 12 matrices of the form (2.1), can be symmetrized by a single matrix (symmetrizer). In addition, they observed that at least two different sym-metrizers exist, based on either the hyperbolic or the parabolic terms.

2.2. The Cauchy and half space problem

The Cauchy problem for (2.1) can be Fourier transformed in all coordinates which yields ˆ

Vt+ i(ωxA + ω¯ yB + ω¯ zC) ˆ¯ V + [ω2xD¯11+ ω2yD¯22+ ω2zD¯33+

+ ωxωy( ¯D12+ ¯D21) + ωyωz( ¯D23+ ¯D32) + ωzωx( ¯D13+ ¯D31)] ˆV = 0, (2.3)

where ωx, ωy, ωz are the wave numbers in x, y, z direction respectively. By multiplying

(2.3) with ˆV∗(the complex conjugated ˆV ) from the left, we get

| ˆV |2t+ 2   iωxVˆ iωyVˆ iωzVˆ   ∗  ¯ D11 D¯12 D¯13 ¯ D21 D¯22 D¯23 ¯ D31 D¯32 D¯33   | {z } ¯ D   iωxVˆ iωyVˆ iωzVˆ  = 0. (2.4)

We demand that the second order terms are parabolic (see (2.2),(2.3),(2.4)) ¯

D > 0, D11> 0, D22> 0, D33> 0,

ω2xD11+ ω2yD22+ ωxωy(D12+ D21) > 0,

ω2yD22+ ωz2D33+ ωyωz(D23+ D32) > 0, (2.5)

ω2zD33+ ωx2D11+ ωzωx(D13+ D31) > 0,

for values of ωx, ωy, ωz not vanishing simultaneously, in which case the Cauchy problem

is well-posed.

The upcoming analysis will be performed in the half space Ω : {x > 0, −∞ < y, z < +∞}, with boundary conditions imposed at the plane x = 0. The Fourier transformed equations in y, z, stemming from (2.1) that we will analyze read

ˆ

Vt+ [ ¯A − iωy( ¯D12+ ¯D21) − iωz( ¯D13+ ¯D31)] ˆVx− ¯D11Vˆxx+

+ [i(ωyB + ω¯ zC) + ω¯ y2D¯22+ ω2zD¯33+ ωyωz( ¯D23+ ¯D32)] ˆV = 0. (2.6)

3. Analysis

Our ambition in this paper is to prove that the energy method and the Laplace trans-form method lead to the exact same number of boundary conditions for the system (2.6).

(4)

3.1. The Laplace transform analysis

To determine the number of boundary conditions, the ansatz ˆV = ψeκx+st, with κ = κR+ iωxand s = η + iξ is inserted into (2.6). We find

(sI + [ ¯A − iωy( ¯D12+ ¯D21) − iωz( ¯D13+ ¯D31)]κ − ¯D11κ2+

+ [i(ωyB + ω¯ zC) ˆ¯ V + ωy2D¯22+ ωz2D¯33+ ωyωz( ¯D23+ ¯D32)])ψ = 0. (3.1)

The number of boundary conditions at x = 0 is given by the number of decaying modes with Re(κ) = κR< 0 for Re(s) = η > 0 obtained from

Det(sI + [ ¯A − iωy( ¯D12+ ¯D21) − iωz( ¯D13+ ¯D31)]κ − ¯D11κ2+

+ i(ωyB + ω¯ zC) ˆ¯ V + ωy2D¯22+ ωz2D¯33+ ωyωz( ¯D23+ ¯D32)) = 0. (3.2)

To solve the problem (3.2) for all roots κ poses a significant challenge. (As an example, for the three-dimensional Navier-Stokes, (3.2) is a 9th degree polynomial in κ.)

Luckily, the are simplifying circumstances. The following lemma will come in handy. Lemma 1. The number of roots, κ, with positive and negative real parts is independent of s and the wave numbers ωy, ωz for <(s) > 0.

Proof. First we show that there are no imaginary generalized eigenvalues κ = iωxfor

Re(s) > 0. By inserting κ = iωxinto (3.1), multiplying from the left with ψ∗and adding

the transpose, we find

ψ∗(ηI + ω2xD¯11+ ωy2D¯22+ ω2zD¯33

+ ωxωy( ¯D12+ ¯D21) + ωyωz( ¯D23+ ¯D32) + ωzωx( ¯D13+ ¯D31))ψ = 0.

By (2.5) we can conclude that ψ = 0 for Re(s) = η > 0.

To complete the proof we need that all roots κ are bounded for bounded (s, ωy, ωz).

Then, for (s, ωy, ωz) restricted to any bounded region R, we can find a closed contour

C bounded by the imaginary axis and a curve in Re(κ) > 0 such that no roots can ever intersect C as (s, ωy, ωz) varies in R. That the number of roots (counting multiplicities)

within C is the same for all points in R follows from the continuity of the logarithmic derivative of the characteristic polynomial on C and an application of the argument prin-ciple (Marden 1949, Ch. 1). The boundedness of κ is established by a simple perturbation argument in Lemma 5, whose proof we leave to the appendix.

Remark 2. Lemma 1 simplifies the analysis since we can set ωy= ωz= 0 and choose

any s with Re(s) > 0 to determine the number of roots κ with positive and negative real parts. We can for example choose s with Re(s) → ∞, which often simplifies the algebra.

We end this section by stating that the number of boundary conditions at x = 0 given by the classical Laplace transform analysis is equal to the number of decaying modes with Re(κ) = κR< 0 for Re(s) > 0, obtained from

Det(sI + ¯Aκ − ¯D11κ2) = 0. (3.3)

3.2. The energy analysis

In the energy analysis for determining the number of boundary conditions, equation (2.6) is multiplied with the solution, and integrated from zero to infinity. We find

|| ˆV ||2t+ 2 Z ∞ 0   ˆ Vx iωyVˆ iωzVˆ   ∗  ¯ D11 D¯12 D¯13 ¯ D21 D¯22 D¯23 ¯ D31 D¯32 D¯33     ˆ Vx iωyVˆ iωzVˆ  dx = BT, (3.4)

(5)

4 J. Nordstr¨om and Thomas Hagstrom where BT =  ˆ V ˆ Vx ∗ ¯ A − ¯D11 − ¯D11 0   ˆ V ˆ Vx  . (3.5) In (3.5), the boundary term BT is evaluated at x = 0, and the values at ∞ are ignored. The total number of boundary conditions in the energy method are given by the number of positive eigenvalues λ in

Det  ¯ A − ¯D11 − ¯D11 0  −  λI 0 0 λI  . (3.6)

Remark 3. The total number of positive and negative eigenvalues can be calculated directly by invoking Lemma 4 below and computing a representation of the matrix in the form RTDR where D is diagonal; see, e.g., (Golub & van Loan 1989, Ch. 4).

3.3. The relation between the Laplace and energy method

The question that we will discuss in this paper arises by comparing equation (3.3) for κ and (3.6) for λ. The two equations have similar ingredients (the matrices ¯A and ¯D11) but

are otherwise quite different. However, to the best of our knowledge, these two methods always provide the same number of boundary conditions. The question is: how can that be possible?

To answer that question, let us return to the governing system of equations (2.6) with the simplifying insertion of ωy= ωz= 0, making the problem one-dimensional

Ut+ ¯AUx− ¯D11Uxx= 0. (3.7)

To ease notation, we replaced ˆV with U in (3.7). The matrix structure in (2.2) suggests the block form

U =  U1 U2  , A =¯  A11 A12 A21 A22  , D¯11=  D11 0 0 0  . (3.8)

Remark 4. The simplification to a purely one-dimensional study, is possible due to Lemma 1 in combination with conditions (3.3) and (3.6).

Next we transform (3.7) to first order form by introducing φ = (U1)x. The new set of

governing equations becomes   I 0 0 0 I 0 0 0 0     U1 U2 φ   t +   A11 A12 −D11 A21 A22 0 −D11 0 0     U1 U2 φ   x +   0 0 0 0 0 0 0 0 D11     U1 U2 φ  = 0. (3.9) The energy method applied to (3.9) yields the energy rate

||U ||2 t+ 2 Z ∞ 0 φTD11φdx = BT (3.10) where BT =   U1 U2 φ   T  A11 A12 −D11 A21 A22 0 −D11 0 0   | {z } E   U1 U2 φ  . (3.11)

(6)

at x = 0 is given by the number of positive eigenvalues λ to

Det (E − λI) = 0. (3.12) To formalise the result above we state

Lemma 2. The eigenvalues satisfying (3.12) are identical to the non-zero eigenvalues in (3.6).

Proof. Follows directly from (3.8).

The Laplace transform method applied to (3.9) with the ansatz (U1, U2, φ)T = ψeκx+st

yields the new generalized eigenvalue problem          A11 A12 −D11 A21 A22 0 −D11 0 0   | {z } E κ +   sI 0 0 0 sI 0 0 0 D11          ψ = 0, (3.13)

where the matrix E present in the energy rate (3.11) shows up again. The generalized eigenvalues κ are given by

Det  Eκ +   sI 0 0 0 sI 0 0 0 D11    = 0. (3.14)

Remark 5. The matrix E is the link between the energy and Laplace transform method.

To guarantee uniqueness, we need

Lemma 3. The determinant condition (3.14) leads to the same eigenvalues κ as (3.3).

Proof. Consider (3.14) and recall that adding a row (column) multiplied by a scalar to another row (column) does not change the determinant. By adding the κ times the lower block rows to the upper block rows we find

Det   A11κ + sI − D11κ2 A12κ 0 A21κ A22κ + sI 0 −D11κ 0 D11  = Det  A11κ + sI − D11κ2 A12κ A21κ A22κ + sI  × Det(D11) = 0.

Since Det(D11) 6= 0, inserting the relations (3.8) into (3.3) proves the claim.

To proceed, we also need the following Lemma from Nordstr¨om & Sv¨ard (2005). Lemma 4. Suppose that R is a nonsingular matrix and that P is a real symmetric matrix. Then the number of positive/negative eigenvalues of RTP R is the same as the

number of positive/negative eigenvalues of P .

Proof. Follows directly from Sylvester’s criterion, see Horn & Johnson (1990). We can now prove the main result of this paper.

(7)

6 J. Nordstr¨om and Thomas Hagstrom

Theorem 1. Consider the half plane problem (2.1) in the domain Ω : {x > 0, −∞ < y, z < +∞} with the boundary conditions imposed at the plane x = 0. The energy and Laplace transform methods lead to the same number of boundary conditions at x = 0.

Proof. We will show that i) the total number of non-zero eigenvalues λ in (3.12) and the total number of non-zero generalized eigenvalues κ in (3.14) for Re(s) > 0 are identical, and that ii) the number of positive λ in (3.12) is the same as the number of eigenvalues κ with Re(κ) < 0 for Re(s) > 0. If both i) and ii) hold, Theorem 1 follows since Lemma 2 and 3 show the equivalence between the first and second order form.

Consider the matrix in (3.14). By Lemma 1 we can choose any s with Re(s) > 0 and get the correct number of eigenvalues κ with negative real part. We choose the specific value s = 1. Since Det(D11) 6= 0 and positive definite, we can factorize D11into D

1/2 11 D 1/2 11 and get   sI 0 0 0 sI 0 0 0 D11  → (s = 1) →   I 0 0 0 I 0 0 0 D1/211     I 0 0 0 I 0 0 0 D1/211  = SS. (3.15)

Now we can insert these matrices into (3.14) noting that S is symmetric and find Det(S(S−1ES−1+ κ−1I)S) = Det(S)2Det((S−1)TES−1+ κ−1I). (3.16) The final result for i) above follows from Lemma 4. Also ii) follows since the number of generalized eigenvalues with Re(κ) < 0 is the same as the number of positive eigenvalues λ of E.

4. Conclusions

We investigated the energy method and the Laplace transform method as techniques for determining the number of boundary conditions required for a well posed initial boundary value problem. By reducing the systems of equations to first order form, we proved that both these methods yield the same number of boundary conditions.

Since the Laplace transform method is known to provide necessary and sufficient condi-tions for the well-posedness of initial-boundary value problems, our result shows that the energy method will as well for hyperbolic-parabolic systems. Also, since both methods can be applied pointwise at smooth boundaries, the result extends to this case. Thus the en-ergy method provides an easy-to-apply general approach to specifying enen-ergy-compatible boundary conditions.

REFERENCES

Abarbanel, S. & Gottlieb, D. 1981 Optimal time splitting for two- and three-dimensional Navier-Stokes equations with mixed derivatives. Journal of Computational Physics 41 (1), 1–33.

Engquist, B. & Gustafsson, B. 1987 Steady state computations for wave propagation prob-lems. Mathematics of Computation 49 (179), 39–64.

Eriksson, S. & Nordstr¨om, J. 2017 Exact non-reflecting boundary conditions revisited:

Well-posedness and stability. Foundations of Computational Mathematics 17 (4), 957–986. Golub, G. & van Loan, C. 1989 Matrix Computations. Baltimore: The Johns Hopkins

Uni-versity Press.

Gustafsson, B., Kreiss, H. O. & Oliger, J. 1995 Time Dependent Problems and Difference Methods. Academic Press.

(8)

Gustafsson, B & Sundstr¨om, A 1978 Incompletely Parabolic Problems in Fluid-Dynamics. SIAM Journal on Applied Mathematics 35 (2), 343–357.

Hersh, R. 1963 Mixed problems in several variables. J. Math. Mech. 12.

Hersh, R. 1964 Boundary conditions for equations of evolution. Arch. Rat. Mech. Anal. 16. Horn, R. A. & Johnson, C. R. 1990 Matrix Analysis. Cambridge University Press.

Kreiss, H.O. 1970 Initial boundary value problems for hyperbolic systems. Communications on Pure and Applied Mathematics 23 (3), 277–298.

Kreiss, H. O. & Lorenz, J. 1989 Initial Boundary Value Problems and the Navier-Stokes Equations. John Wiley & Sons.

Lauren, F. & Nordstr¨om, J. 2018 Practical inlet boundary conditions for internal flow

cal-culations. Computers and Fluids 175, 159–166.

Marden, M. 1949 The Geometry of the Zeros of a Polynomial in a Complex Variable. New York: American Mathematical Society.

Nordstr¨om, J. 1989 The influence of open boundary conditions on the convergence to steady

state for the Navier-Stokes equations. Journal of Computational Physics 85 (1), 210–244.

Nordstr¨om, J. 2017 A Roadmap to Well Posed and Stable Problems in Computational Physics.

Journal of Scientific Computing 71 (1), 365–385.

Nordstr¨om, J. & Gustafsson, R. 2003 High order finite difference approximations of

elec-tromagnetic wave propagation close to material discontinuities. Journal of Scientific Com-puting 18 (2), 215–234.

Nordstr¨om, J., Mattsson, K. & Swanson, C. 2007 Boundary conditions for a divergence

free velocity-pressure formulation of the navier-stokes equations. Journal of Computational Physics 225 (1), 874–890.

Nordstr¨om, J. & Sv¨ard, M. 2005 Well-posed boundary conditions for the Navier-Stokes

equations. SIAM Journal on Numerical Analysis 43, 1231–1255.

Sakamoto, R. 1982 Hyperbolic Boundary Value Problems. Cambridge: Cambridge University Press.

Strikwerda, J.C. 1977 Initial boundary value problems for incompletely parabolic systems. Communications on Pure and Applied Mathematics 30 (6), 797–822.

5. Appendix

Lemma 5. For bounded (s, ωy, ωz) and Re(s) = η > 0, the roots κ of (3.2) are bounded.

Proof. Rewrite the eigenvalue problem (3.1) in block form according to the dimensions of D11 and recall that ¯A, ¯B and ¯C are symmetric:

 ˆP11+ ˆQ11κ + sI − D11κ2 ˆV 1+ iωyB¯12+ iωzC¯12+ ¯A12κ ˆ V2= 0, (5.1) iωyB¯12T + iωzC¯12T + ¯AT12κ ˆ V1+ iωyB¯22+ iωzC¯22+ ¯A22κ + sI ˆ V2= 0, (5.2)

where we have introduced ˆ

P11= sI + i ωyB¯11+ ωzC¯11 + ω2yD22+ ω2zD33+ ωyωz(D23+ D32) ,

ˆ

Q11= ¯A11− iωy(D12+ D21) − iωz(D13+ D31). (5.3)

Now suppose that a uniformly bounded sequence of (s, ωy, ωz), Re(s) = η > 0 exists

with |κ| → ∞. Since D11is nonsingular, equation (5.1) then implies ˆV1→ 0. This makes

the largest term in equation (5.2) to be κ ¯A22Vˆ2, so we must have ¯A22Vˆ2 → 0. Since we

are looking for nonzero solutions, such a scenario is only possible if ¯A22 is singular with

ˆ

V2 approaching one of its null vectors, v0. Returning to (5.1) and collecting the largest

terms we find

(9)

8 J. Nordstr¨om and Thomas Hagstrom from which we conclude

ˆ

V1∼ κ−1D−111A¯12v0. (5.5)

Now set ˆV2∼ v0+ κ−1w0 and consider O(1) terms in (5.2):

¯

AT12D11−1A¯12v0+ iωyB¯22v0+ iωzC¯22v0+ ¯A22w0+ sv0= 0. (5.6)

Multiplying by vT

0 and taking the real part yields:

vT0A¯T12D−111A¯12v0+ ηvT0v0= 0, (5.7)

where we have used the symmetry of ¯A22, ¯B22 and ¯C22and the fact that vT0A¯22w0= 0.

Since η > 0 and the first term in (5.7) is nonnegative, we have reached a contradiction. Thus κ is bounded.

(10)

References

Related documents

This first im- plementation provided an ontology and a knowledge base (KB) for storing a set of objects and properties, and spatial relations between those objects. The KB

Användningen av Facebook verkar vara konfliktfylld, där personer å ena sidan inte vill att det ska spela roll, men också beskriver hur det faktiskt gör det. Varför kan

In addressing this call for research, the purpose of this paper is to contribute to the field of entrepreneurship in established organizations by the study of an

Livsstilsfaktorer som också beskrivs öka risken för försämrad näringsstatus hos äldre, är att leva ensam, ha långvariga alkoholproblem samt låg kroppsvikt innan sjukdom

”Jag har uppsyn i Knivsta och Uppsala, så ringer det någon från Knivsta 22 eller 23 mil bort och säger att det står tre killar på taket utan skydd och ber mig att komma och

Att delprojektet genomfördes vid SMP Svensk Maskinprovning AB (SMP) avgjordes av att SMP förfogar över motorbromsar som kan styras så att transienta förlopp kan återskapas och

The aim was to evaluate from a stakeholders view point, the feasibility of utilising mobile phone technology in the Kenya’s reproductive health sector in Nakuru Provincial

I gruppen med måttlig sannolikhet för instabil angina hade bara 3% av patienterna akut myokard infarkt (AMI). Av dessa patienter behövde bara 37% invasiv