• Keine Ergebnisse gefunden

Analyzing the stability behaviour of DAE solutions and their approximations

N/A
N/A
Protected

Academic year: 2022

Aktie "Analyzing the stability behaviour of DAE solutions and their approximations"

Copied!
36
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Analyzing the stability behaviour of DAE solutions and their approximations

Roswitha Marz

Antonio R. Rodrguez Santiesteban and Humboldt-University Berlin

Institute of Mathematics Unter den Linden 6 D-10099 Berlin, Germany

1

(2)

Contents

1 Introduction 3

2 Analysis of linear index-2 DAE's 6

2.1 Fundamentals . . . 6

2.2 Solvability and decoupling . . . 8

2.3 Index reduction by dierentiation . . . 12

2.4 Contractivity . . . 16

3 Analyzing numerical integration methods for linear index-2 DAEs 20

3.1 BDF . . . 20

3.2 Runge-Kutta methods . . . 24

4 Two illustrating examples 30

2

(3)

1 Introduction

Usually, lower-index dierential-algebraic equations (DAEs)

f(x0(t) x(t) t) = 0 (1.1)

are integrated by numerical algorithms that have been developed originally for reg- ular ordinary dierential equations (ODEs). There are numerous papers justifying this by convergence proofs, order results etc. (cf. 1], 2], 3]). In particular, the backward dierentiation formula (BDF)

f

0

@1 h

k

X

j=0jxn;j xn tn

1

A= 0 (1.2)

but also implicit Runge-Kutta methods (IRK) xn=xn;1+hXs

j=1jXnj0 (1.3 a)

f

0

@Xni0 xn;1+hXs

j=1ijXnj0 tn;1+cih

1

A= 0 i= 1 ::: s (1.3 b) are used in applications with really great success. Step by step each method provides numerical approximations xn of the true solution values x(tn) tn=tn;1+h. The convergence results mentioned above concern the behaviour of the global error on a compact interval, say t0 T], as the stepsize of the discretizationt0 < t1 < <

tn=T tends to zero.

However, what do we know about the error x(tn);xn if the stepsize h is constant and n ! 1, that is, tn ! 1? When integrating regular ODEs numerically, one tries to carefully match the dynamics of the numerical algorithm with the dynamical behaviour of the true solution. How could this be done in case of DAEs?

Classical linear stability theory is concerned with the analysis of approximating the so-called scalar linear test equation

z0(t) =z(t) Re <0: (1.4) Basic notions like the region of absolute stability, error growth function etc., rely on (1.4) and apply, via similarity transforms, to constant coecient regular ODEs

x0(t);Wx(t) = 0: (1.5)

Considering DAEs, the respective constant coecient system

Ax0(t) +Bx(t) = 0 (1.6)

3

(4)

has a singular leading coecient matrixA, but a regular matrix pencilfA Bg, i. e., det(A+B)60. There are nonsingular matricesE F that transform (1.6) into its so-called Kronecker normal form

I J

!

x~0(t) +

;W I

!

x~(t) = 0 (1.7)

where EAF =

I J

!

EBF =

;W I

!

x~ = F;1x and J is a nilpotent block, ind(J) = indfA Bg.

Obviously, the solution of the homogeneous equation (1.6) consists of its "dynamical part" only, i. e.,

x(t) =Fx~(t) = F

u~(t) 0

!

where ~u(t) solves the regular ODE ~u0(t) =Wu~(t).

On the other hand, we may apply the same transformations E F to decouple the BDF applied to (1.6), that is,

A1h

k

X

j=0jxn;j+Bxn = 0 to

I J

!1 h

k

X

j=0jx~n;j+

;W I

!

x~n = 0:

Starting with consistent values x0 xk;1 (such that F;1xj =

u~j

0

!

j = 0 k) we obtain xn=Fx~n =F

u~n

0

!

where ~un is given by the BDF applied to the regular ODE ~u0(t) =Wu~(t), i. e., h1

k

X

j=0ju~n;j =Wu~n:

In the same way we may proceed with Runge-Kutta methods. Obviously, the rich world of classical stability theory applies to constant coecient DAEs (1.6) via transformation into Kronecker normal form plus similarity transform.

However, unfortunately, the homogeneous coecient DAE (1.6) does not play a similar role in the DAE analysis as equation (1.5) does in the regular ODE-theory.

DAEs represent a much more complex class of problems. The constant coecient 4

(5)

case (1.6) is only a very poor model, which does not reect important geometric features of DAEs at all.

Positive results concerning long term integrations of index-1 DAEs are reported in 4]. There, the leading nullspace is supposed to remain invariant. Moreover, in 4]

it is stressed that varying subspaces may have its eect.

In 5], the following small example was discussed to show the bad eect of a rotating nullspace. The linear index-1 DAE

;1 t

0 0

!

x0(t) +

;1 t

;1 t;1

!

x(t) = 0 (1.8)

with real parameters 6= 0 6= 1 has the solutions x1(t) = (;1);1(1;t)x2(t)

x2(t) = exp((;)t)x2(0):

The nullspace of the leading coecient matrix is N(t) :=nz 2IR2 : (;1)z1+tz2 = 0o:

For 6= 0, this nullspace varies with t. Applying the backward Euler method to (1.8) yields

xn1 = (;1);1(1;tn)xn2

xn2 = 1+1+hhxn;12:

Obviously, it holds that xn ! 0 (n ! 1) if and only if j1 +hj < j1 +hj. For h = ;1 the method is not feasible at all. Further, there is a large region on the (h h)-plane where the true solution x(tn) vanishes asymptotically, but the numerical approximation grows unboundedly.

Note that the spectrum of the pencil fA(t) B(t)gis time-invariant, namely (A(t) B(t)) := f2IC : det(A(t) +B(t)) = 0g=f;g:

If we put = 0, we obtain again a constant coecient DAE and everything is ne as expected before.

One could think that the strange asymptotic behaviour is due to the rotations of the nullspace of the leading coecient A(t) or the leading partial Jacobian fx00(x0 x t) in (1.1). Since this matrix has a constant nullspace or is constant itself in most applications, we can perhaps consider equation (1.8) to be just an academic example.

However, similar phenomena of a "wrong numerical dynamical behaviour" may arise even in the case of a constant leading coecient matrix. Such phenomena are discussed in 6], where linear index-2 DAEs are considered.

5

(6)

Roughly speaking, in 6] it is shown that the numerical approximations t the dy- namical behaviour of the true solution well, supposed both the characteristic sub- spaces N1(t) andS1(t) do not vary with t in fact.

The aim of the present paper is to show that the same invariance condition forS1(t) will do. This weaker sucient condition ensures also an appropriate reection of the exponential decay. Since e.g. in circuit simulation S1(t) is often invariant but N1(t) is not, our new approach is of particular interest for those applications.

This paper is organized as follows. In Section 2 the related analytical background is given. Useful analytical decoupling and reduction techniques are discussed. As a by product, Theorem 2.2 provides new solvability results with respect to weaker smoothness demands. Also contractivity is now discussed under those weaker smooth- ness conditions.

In Section 3 the analytical decoupling and reduction techniques are applied to BDF and IRK methods, to prove that the time invariance of the subspace S1(t) will do in fact.

Finally we demonstrate by two characteristic examples that are even in Hessenberg form, how a rotating subspace S1(t) eects the dynamic reection.

2 Analysis of linear index-2 DAE's

2.1 Fundamentals

Consider the linear equation

A(t)x0(t) +B(t)x(t) =q(t) t2J := t0 1) (2.1) where the coecients are continuous and the matrix A(t) 2 L(IRm) is singular for all t2J, but has constant rankr. Introduce the basic subspaces

N(t) := kerA(t)

S(t) := fz 2IRm :B(t)z 2 imA(t)g:

Obviously, each solution of the homogeneous equation satises x(t)2S(t) t2J:

In the following, we assumeN(t) to be spanned bym;r continuously dierentiable base functions. Then, there is a C1 matrix functionQ : J !L(IRm) that projects IRm pointwise onto N(t) i. e., Q(t)2 =Q(t) im Q(t) =N(t) t2J.

In what follows, Q denotes such a C1 projector function onto N, and P :=I;Q. Recall (e.g. 2]) the equation

A(t)f(Px)0(t);P0(t)x(t)g+B(t)x(t) =q(t) t 2J (2.2) 6

(7)

to be the precise formulation of (2.1). We are looking for continuous functions x( ) :J !IRm that have continuously dierentiable parts (Px)( ), and that satisfy (2.2) pointwisely. Denote this function space by

CN1 :=ny2C(J IRm) :Py2C1(J IRm)o:

Next, we introduce further subspaces which are relevant for index-2 DAEs 2], namely

N1(t) := kerA1(t)

S1(t) := fz2IRm :B(t)P(t)z 2 imA1(t)g A1(t) := A(t) + (B(t);A(t)P0(t))Q(t) t2J:

Denition:

The DAE (2.1) is said to be index-2 tractable if dim(N(t)\S(t)) = >0 N1(t)\S1(t) =f0g t2J.

Recall that index-2 tractability generalizes the case of Kronecker index 2. Let Q1(t) denote the projector onto N1(t) alongS1(t) t2J.

Due to 4], Lemma A.13, the matrix G2(t) :=A1(t) +B(t)P(t)Q1(t)

remains nonsingular now and it holds that Q1(t) = Q1(t)G2(t);1B(t)P(t) on J. Since we may represent

kerA1(t) = (I;P(t)A(t)+(B(t);A(t)P0(t))Q(t))(N(t)\S(t)) we know N1(t) to have the same constant dimension as N(t)\S(t).

In the following, we drop the argument t if possible.

By straightforward calculations we prove the relations

G;12 A =P1P G;12 B =G;12 BPP1+Q1+Q+P1PP0Q:

Hence, scaling equation (2.1) by G;12 leads to

P1P((Px)0;P0x) +G;12 BPP1x+Q1x+Qx+P1PP0Qx=G;12 q: (2.3) Multiplying (2.3) by PP1 QP1 and Q1, respectively, and carrying out simple com- putations we obtain the decoupled system

PP1(Px)0+PP1G;12 BPP1x = PP1G;12 q (2.4)

;QQ1(Px)0+QP1G;12 BPP1x+Qx = QP1G;12 q (2.5)

Q1x = Q1G;12 q: (2.6)

7

(8)

The System (2.4), (2.5), (2.6) provides the basic idea of what the solutions of the DAE (2.1) look like in case of PP1 PQ1 and PQ1G;12 q beingC1. Then (2.4), (2.5) may be rewritten as

(PP1x)0 ;(PP1)0PP1x+PP1G;12 BPP1x=PP1G;12 q+ (PP1)0PQ1G;12 q (2.4') Qx=;(QP1G;12 B+QQ1(PQ1)0)PP1x + QQ1(PQ1G;12 q)0+QP1G;12 q

; QQ1(PQ1)0PQ1G;12 q (2.5') and the solution is composed by x = PP1x+PQ1x+Qx where the component PP1xsolves the inherent regular ODE (2.4'),Qx and PQ1x are given by (2.5') and (2.6), respectively. (cf. 2]).

In the next section, we will generalize the solvability results given in 8] in the sense that we will do with lower smoothness. The new inherent regular ODE, which uses a dierent projector instead of PP1, permits to obtain new insights into the asymptotic stability behaviour of numerical approximations (Section 3).

Note that, by construction, P(t)P1(t) projects onto P(t)S1(t) along N(t)N1(t), i. e., it is related to the decomposition

IRm =P(t)S1(t)N(t)N1(t): (2.7)

2.2 Solvability and decoupling

Now we also use the orthoprojectorV(t), which projects alongS1(t). Then,I;V(t) is the orthoprojector ontoS1(t). Because ofN(t) S1(t), it holds thatV(t)Q(t) = 0, and

(t) :=P(t)(I;V(t))

is a projector, too. Obviously, we have im (t) = P(t)S1(t). Since the two pro- jectors (t) and P(t)P1(t) have the same image space P(t)S1(t), it holds that PP1 = PP1 =PP1.

Additionally, we denote the orthoprojector onto im A1(t) by I;W(t).

As we will realize below, the projection W extricates exactly that derivative free part of equations that has to be dierentiated once when reducing the index.

The following example of a Hessenberg form DAE is to make the meaning of the dierent projectors more transparent.

Example:

Given a Hessenberg form DAE of size 2 that contains n1 equations with derivatives and n2 derivative-free ones

x01+ B11x1+B12x2 = q1 B21x1 = q2

)

B21B12 nonsingular.

8

(9)

Here we have a constant nullspace N N =fz :z1 = 0g further Q=

0 0 0 I

!

A1 =

I B12 0 0

!

W =

0 0 0 I

!

S(t) =S1(t) =fz :B21(t)z1 = 0g S(t)\N =N N1(t) =fz:z1 +B12(t)z2 = 0g

N1(t)\S1(t) =fz :z1+B12(t)z2 = 0 B21(t)z1 = 0g=f0g due to the nonsingularity of B21B12.

The projector PP1 is given by PP1 =

I;L 0

0 0

!

where L =B12(B21B12);1B21 is also a projector (L(t) projects IRn1 onto im B12(t) along kerB21(t)). The orthoprojector along S1(t) is now

V(t) =

B21(t)+B21(t) 0

0 0

!

and (t) =

I;B21(t)+B21(t) 0

0 0

!

im (t) = P(t)S1(t) =fz :B21(t)z1 = 0 z2 = 0g

= kerB21(t)f0g:

It is well-known that, in Hessenberg systems, all derivative-free equation should be dierentiated when deriving a solution via a reduction step.

Consequently, it seems to be more natural to assume B21 2 C1 (hence, V 2C1) instead of L2C1 (i. e., PP1 2C1).

2

Next, we collect some nice properties of our projectors and matrices to be used below:

1) im A1 = kerW im A1 = kerG2PQ1G;12 , hence there are two projectors W and G2PQ1G;12 along imA1. Therefore

W =WG2PQ1G;12 G2PQ1G;12 =G2PQ1G;12 W:

9

(10)

2) FromWB =WG2PQ1G;12 B =WG2PQ1G;12 BPQ1 =WBPQ1 and PQ1G;12 WB =PQ1 it follows that

kerWB = kerPQ1 = kerQ1 =S1 imWB = im W:

3) V = (WB)+WB W = (I;A1A+1) = (PQ1G;12 )+PQ1G;12 WB = (PQ1G;12 )+PQ1:

4) V =V PQ1 PQ1 =PQ1V V =V P V =V Q1 V G;12 = (WB)+WBG;12 = (WB)+WBPQ1G;12

= (WB+WG2PQ1G;12 BPQ1G;12 = (WB)+WG2PQ1G;12 = (WB)+W. The third property indicates that, for continuously dierentiablePQ1G;12 andPQ1, we also have continuously dierentiable W and V. However, as we could realize by means of the special Hessenberg form DAE above, the opposite is not true. In this sense, Theorem 2.2 below generalizes the related results of 8].

To get a better insight we reconsider the decoupled form (2.4), (2.5), (2.6) of the DAE (2.1) and try to reformulate it in such a way that (2.4) returns into a regular ODE for the solution component x. Now, we do not assume PP1 PQ1 to be from C1 as we did in order to obtain (2.4'), (2.5'), but we allow only P and V to be continuously dierentiable now.

Equation (2.6) yields immediately

V x= (WB)+Wq: (2.8)

To realize what equation (2.4) looks like in more detail we derive PP1(Px)0 = PP1( +PV)(Px)0 = (Px)0+PP1V(Px)0

= (x)0;0Px+PP1(V x)0;PP1V0Px:

Consequently, (2.4) can be rewritten as

(x)0;0(x+ PV x);PP1V0(x+PV x) +PP1(V x)0

+ PP1G;12 BPP1(x+PV x) =PP1G;12 q: (2.9) Analogously, with QQ1(Px)0 = QQ1V(Px)0 = QQ1(V x)0 ; QQ1V0(x+ PV x), equation (2.5) can be reformulated:

;QQ1(V x)0+QQ1V0(x+PV x)+QP1G;12 BPP1(x+PV x)+Qx=QP1G;12 q:(2.10)

10

(11)

If x2CN1 solves the DAE (2.1), then (WB)+Wq=V x=V Pxbelongs toC1 since V and Px do so. We are allowed to replace V x and (V x)0 in (2.9) and (2.10) by means of (2.8). This gives rise to consider the ODE

u0;0u ;PP1V0u+PP1G;12 BPP1u

= 0P(WB)+Wq+PP1V0P(WB)+Wq;PP1G;12 BPP1(WB)+Wq

;PP1((WB)+Wq)0+PP1G;12 q (2.11) and to call it an inherent regular ODE. If we multiply equation (2.11) by (I;), we obtain

(I ;)u0 ;(I ;)0u= 0 hence

((I;)u)0+ 0(I;)u= 0:

It results that if u 2C1 solves (2.11), then the function ! = (I;)u satises the homogeneous regular ODE !0+0!= 0. If, additionally,!(t) = (I;(t))u(t) = 0, then !(t) vanishes identically, i. e., u(t) = (t)u(t) t 2 J. This proves the following assertion.

Lemma 2.1

The subspace im (t) IRm represents an invariant subspace of the regular inherent ODE (2.11).

As far as the homogeneous case q = 0 is concerned, we know from (2.8) that each solution of Ax0 +Bx = 0 has a trivial component V x = 0. Using (2.9), (2.10) and taking into account the inherent regular ODE we may express each solution as x = x +Qx = canu, where u solves (2.11) as well as the initial condition u(to)2 im (t0), and

can:=KPP1 K := (I;QP1G;12 BPP1;QQ1V0PP1):

The matrix function K is nonsingular, thus ker can(t) = kerP(t)P1(t) t 2J. It is easily checked that can is also a projector. canis said to be the canonical projector for the index-2 case. It projects onto the solution space of the homogeneous DAE along N N1, as we will see below.

Theorem 2.2

Given an index-2 DAE (2.1) with continuously dierentiable sub- spaces N S1 and im A1. Additionally, let WB be C1 and q2fp2C:Wp2C1g.

(i) Then, the IVP for (2.1) with the initial condition (t0)(x(t0);x0) = 0 x0 2IRm, has exactly one CN1-solution.

(ii) Exactly one solution of the homogeneous equation at t0 passes through each x0 2im can(t0) .

11

(12)

Proof:

(i) Denote by u 2 C1 the solution of the regular ODE (2.12) that satises the initial conditionu(t0) = (t0)x0. Then we compose the functionx=u+v+w v := (WB)+Wq2C1

w:=QP1G;12 q+QQ1v0;QQ1V0(u+Pv);QP1G;12 BPP1(u+Pv): Due to this construction we have v =V v w=Qw Px=u+Pv 2C1 x2C (t0)x(t0) =u(t0) = (t0)x0.

Finally, straightforward checking shows thatxsatises the DAE (2.1), indeed.

(ii) Solve the special IVP for q = 0 and x0 := x0 = can(t0)x0. Its solution is x= canu, and we have, in particular,

x(t0) = can(t0)u(t0) = can(t0)(t0)x0

= can(t0)(t0)can(t0)x0 = can(t0)(t0)(PP1)(t0)x0

= can(t0)(PP1)(t0)x0

= can(t0)x0 =x0:

2

Corollary:

On each compact interval t0 T], the perturbation index of an index-2 tractable DAE is two.

Proof:

For each T > t0, there is a constantKT such that the estimation

kxkT KT

n

j(t0)x0j+kqkT +k(V q)0kT

o (2.12)

holds for all IVP solutions with x0 2IRm q2C V q2C1, where

kykT := maxfjy(t)j:t2t0 T]g.

2

Example

: For the Hessenberg form DAE x01+ B11x1+B12x2 = q1

B21x1 = q2

)

the conditions of Theorem(2.2) are satised if q1 2C q2 2C1 and B212C1. The initial condition reads in detail

0 = (t0)(x(t0);x0) =

(I;B21(t0)+B21(t0))(x1(t0);x01) 0

!

:

2.3 Index reduction by dierentiation

Considering the DAE (2.1) we observe immediately that each solution has always to proceed in the set

M(t) := fz 2IRm :B(t)z;q(t)2 imA(t)g=

fz 2IRm : (I;W(t))(B(t)z;q(t))2 imA(t) W(t)(B(t)z;q(t)) = 0g 12

(13)

which is called a constraint manifold. Under the conditions of Theorem 2.2, the part of equations described by

WBx=Wq

can be dierentiated to obtain

WBx0+ (WB)0x= (Wq)0:

Here, WBx0 stands forWBf(Px)0;P0xg:On the other hand, due to (2.1) we may express Px0 := (Px)0 ;P0x=PA+(q;Bx): In consequence, the relation

WBA+(q;Bx) +W(WB)0x=W(Wq)0

has to be satised additionally, i. e., the solution has also to proceed in the so-called hidden constraint manifold

H(t) := fz 2IRm : (;W(t)B(t)A(t)+B(t) +W(WB)0(t))z

=W(Wq)0(t);W(t)B(t)A(t)+q(t)g: As we will see below, the set

M

1(t) := M(t)\H(t)

is characteristic of the DAE (2.1). M1 contains all solutions (for given q), but it is lled by those solutions. One could callM1(t) the state manifold. For the case of a homogeneous equation, S(t) = M(t) is trivially given. Moreover, there is a closed relationship between our subspace S1(t) and the state manifoldM1(t).

Lemma 2.3

For a homogeneous index-2 tractable DAE (2.1) with continuously dif- ferentiable N S1 im A1 WB it holds that

M

1(t)jq=0 =nz 2S1(t) :Q(t)z =;((QP1G;12 B+QQ1G;12 (WB)0)PP1)(t)zo and

P(t)M1(t)jq=0 =P(t)S1(t): (2.13)

Proof:

The second relation is a simple consequence of the representation of M1(t), which we want to realize now. We drop the argument t.

z 2M1 means by denition:

;WBA+Bz+W(WB)0z = 0 Bz+Aw= 0 for a certain w=Pw. This is equivalent to

WBw+W(WB)0z = 0 w=;PA+Bz PQ1z = 0 Qz;QQ1w+QP1G;12 BPP1z+QP1PP0Qz = 0, and to

13

(14)

PQ1z = 0 w=;PA+Bz 0 =WBw+W(WB)0PP1z+W(WB)0Qz = WBw+W(WB)0PP1z+WBP0Qz

Qz;QQ1w;QQ1P0Qz+QP1G;12 BPP1z = 0:

Because of QQ1G;12 WB = QQ1G;12 B = QQ1 QQ1G;12 W = QQ1G;12 , we nd z 2M1 to be characterized by

PQ1z = 0 w =;PA+Bz

QQ1w+QQ1P0Qz+QQ1G;12 (WB)0PP1z= 0 Qz = QQ1w+QQ1P0Qz;QP1G;12 BPP1z

= ;QQ1P0Qz+QQ1P0Qz;QQ1G;12 (WB)0PP1z;QP1G;12 BPP1z

= ;QQ1G;12 (WB)0PP1z;QP1G;12 BPP1z

2

Note that the linear space M1(t)jq=0 represents the tangent space for M1(t). Of course, this is much more important for nonlinear problems. Lemma 2.3 shows P(t)S1(t) to be a kind of practical substitute of the tangent space. It should be stressed once more that we are looking forCN1 solutions and that there is no natural need for Q-components of the elements of the tangent spaces.

Next, rewrite the DAE (2.1) as

Ax0+ (I;W)(Bx;q) +W(Bx;q) = 0

and replace the part W(Bx;q) = 0 by its dierentiated form WfWBx0+ (WB)0x;(Wq)0g= 0

to compose the new DAE

(A+WB)x0+ ((I;W)B+W(WB)0)x= (I ;W)q+W(Wq)0: (2.14) Denote ~A := A +WB B~ := (I ; W)B +W(WB)0 and consider (2.14) in more detail. Since (A(t) +W(t)B(t))z = 0 decomposes into Az = 0 WBz = 0, we have ker ~A(t) kerA(t) = N(t). In consequence, the solutions of (2.15) belong to the same class CN1 as the solutions of (2.1). Further, we have

S~(t) =fz 2IRm: ~B(t)z 2 im ~A(t)g

=fz 2IRm: (I;W(t))B(t)z 2 imA(t)

W(t)(WB)0(t)z =W(t)B(t)A(t)+(I;W(t))B(t)zg

f

M(t) =fz 2IRm : (I ;W(t)(B(t)z;q(t))2 im A(t)

W(t)(WB)0(t)z;W(t)(Wq)0(t) = W(t)B(t)A(t)+(I;W(t))(B(t)z;q(t))g

f

M(t)\fz2IRm :W(t)(B(t)z;q(t)) = 0g=M1(t): 14

(15)

Theorem 2.4

Given the conditions of Theorem 2.2.

(i) Then equation (2.14) is index-1 tractable.

(ii) Exactly one solution at t 2 J passes through each x 2Mf(t) . (iii) M1(t) and M(t) form invariant subsets for (2.14).

Proof:

(i) We check the nonsingularity of the matrix ~G1(t) 2L(IRm) ~G1 :=A+WB + ((I;W)B+W(WB)0)Q:Again, we drop the t-argument. ~G1z = 0 yields Az+(I;W)BQz = 0 WBz =;W(WB)0Qz =;WBP0Qz, hencePQ1z =

;PQ1P0Qz (A+BQ)z = 0.

Because of (A+BQ)(I;PP0Q) =A1 A1(I+PP0Q) =A+BQ, the element (I +PP0Q)z = ~z belongs to kerA1 =N1 , i. e., ~z =Q1z:~ On the other hand, PQ1z +PQ1P0Qz = 0 implies PQ1(I +PP0Q)z = 0, i. e., PQ1z~ = 0 and nally ~z = 0 z = 0:

(ii) Due to the index-1 property, this assertion is now given by the index-1 results, e. g. in 2].

(iii) Let x denote a solution of the index-1 tractable DAE (2.14) and x(t)2M(t): Then, (2.14) gives

Ax0+ (I;W)Bx= (I;W)q WBx0+W(WB)0x=W(Wq)0: Consider the function :=W(Bx;q). Derive

0 = (WB)0x+WBx0 ;(Wq)0

= (WB)0x;W(WB)0x+W(Wq)0;(Wq)0

= (I;W)(WB)0x;(I;W)(Wq)0

= ;(I;W)0(WBx;Wq) =W0:

Since x(t)2 M(t) implies(t) =W(t)(B(t)x(t);q(t)) = 0, the func- tion vanishes identically, hence Ax0+Bx=q is satised.

Therefore, x(t) 2 M(t) implies x(t) 2 M(t) but also x(t) 2 M1(t) for all t 2J.

2

15

(16)

Corollary:

The index-2 tractable DAE has dierentiation index two.

Example:

For the Hessenberg system x01 + B11x1 + B12x2 = q1

B21x1 = q2

)

we have now

M(t) = fz :B21(t)z1 ;q2(t) = 0g

H(t) = fz :;B21(t)B11(t)z1;B21(t)B12(t)z2+ +B210 (t)z1 =q02(t);B21(t)q1(t)g

M

1(t) = fz :B21(t)z1 =q2(t) (B21(t)B12(t))z2 =

B210 (t)z1 ;B21(t)B11(t)z1;q20(t) +B21(t)q1(t)g

~

M(t) = fz :B210 (t)z1 ;q20(t) =B21(t)(B11(t)z1 +B12(t)z2;q1(t))g

~

M(t) \ fz :B21(t)z1 =q2(t)g=M1(t): The index-reduced equation (2.15) is, as expected,

x01 + B11x1 + B12x2 = q1 B21x01 + B210 x1 = q02

)

2.4 Contractivity

Now we turn to the asymptotic behaviour of the solutions of the homogeneous equations

Ax0+Bx= 0: (2.15)

As we know, the solutions may be represented by x=KPP1u= canu

where u solves the inherent regular ODE

u0 = 0u+PP1V0u;PP1G;12 BPP1u (2.16) u(t0)2 im (t0):

Recall once more that im (t) is an invariant subspace of the ODE (2.16), i. e., u(t) = (t)u(t) t2J:

16

(17)

Obviously, the stability behaviour of x is mainly governed by the dynamics of the inherent regular ODE, that is, by u. Hence, it would be nice to formulate criteria on how the ow of regular ODE behaves within an given invariant subspace.

The matrix coecient of the inherent regular ODE (2.16) has to be expected to be- come singular. In particular, for constant subspacesN andS1, it is simply the matrix PP1G;12 BPP1, which has at least the nullspaceNN1 of dimensionm;r+ 2.

The standard approaches to estimate how the solutions grow usually relate to all solutions of (2.16), included those solutions that do not start in im (t0). No ex- ponential decay can be realized in this way. The standard techniques are somewhat coarse. Hence, we try to improve them by relating all things to the invariant sub- spaces.

For a given subspace U IRm we introduce the matrix-semi-norm

kGkU := max

(

jGzj

jzj :z 6= 0 z 2U

)

such that jGzjkGkUjzj holds for allz 2U.

Then, we dene the logarithmic matrix norm relative to the subspace U by U(G) := limh

!0

h1(kI+hGkU ;1): (2.17)

This logarithmic norm relative toU is well-dened and has similar properties as the standard version of Dahlquist for U = IRm (cf.(9]).In fact this generalization is in fact very natural and straightforward. On the other hand, it is worth mentioning that U(G) appears to be a special case of the logarithmic norm proposed in 10]

for matrix pencils.

Lemma 2.5

Given a regular linear ODE

u0(t) +M(t)u(t) = 0 t 2J = t0 1)

which has the invariant subspace U(t) IRm t2J. Let the function (t) :=Zt

t0 U(s)(;M(s))ds t2J be well-dened, i. e., the integrals do exist.

Then, for all ODE solutions starting with initial values u(t0)2U(t0), the inequality

ju(t)je(t)ju(t0)j tt0 is valid.

17

(18)

Proof:

This proof follows the lines of the standard theory 9], too. Given a solution with u(t0)2U(t0) we have, u(t)2U(t) for allt2t0 1).

Denote m(t) := ju(t)j t 2t0 1), and derive

m(t+h) = ju(t) +hu0(t) +o(h)j=j(I;hM(t))u(t) +o(h)j

k(I;hM(t)kU(t)ju(t)j+o(h) thus

h1(m(t+h);m(t)) 1

h(kI;hM(t)kU(t);1)m(t);o(h0): Letting h!0, we obtain

D+m(t)U(t)(;M(t))m(t) t 2t0 1)

where D+m(t) denotes the respective Dini derivative. By Peano's Lemma it follows immediately that

m(t)e(t)m(t0):

2

Theorem 2.6

Let (2.15) be index-2 tractable with N S1 and WB from C1. Then, the estimation

jx(t)jkcan(t)ke(t)j(t0)x(t0)j tt0 (2.18) with (t) := tRt

0

(PS1)()(;M())d t t0, holds true for each solution, provided that (t) is well-dened. Here,

M :=;0;PP1V0+PP1G;12 BPP1

is the coecient matrix of the inherent regular ODE (2.16).

Proof:

Applying Lemma 2.5 to (2.16), which has the invariant subspace im (t) = P(t)S1(t), yields ju(t)je(t)ju(t0)j tt0, provided that u(t0)2 im (t0).

The DAE solutions have the representationx(t) = can(t)u(t), andu(t0) = (t0)x(t0).

2

18

(19)

Corollary:

If there is a > 0 such that (PS1)(t)(;M(t)) ; t t0, all DAE solutions satisfy the inequality

jx(t)jkcan(t)ke;(t;t0)j(t0)x(t0j t t0: (2.19) Hence, if can(t) grows "moderately" like polynomials or is bounded by

kcan(t)kCet tt0, with < , then the solutions decrease exponentially.

To realize whether numerical approximations reect the stability behaviour of the true solution well, one often uses the notion of contractivity and so-called one-sided Lipschitz conditions (cf. (9]). This approach applies also to DAEs provided that we relate the things again to our invariant subspaces.

Standard numerical integration methods applied to index-2 DAE are expected to work well if the basic nullspace N is constant. However, it is well-known that these methods may fail in case this nullspace rotates with time. This is why we suppose P0 = 0 now.

If the vector norm used to dene the logarithmic norm is related to an inner product, it holds that

hGz ziU(G)jzj2 for allz 2U:

Denition:

An index-2 tractable DAE with constant P is said to be contractive if there are an inner product h i and a constant >0 such that the inequality

hy Pxi;jPxj2 (2.20)

is valid for all y x2IRm t2t0 1), with

A(t)y+B(t)x= 0 Qy= 0 V(t)y=V(t)0(t)x: (2.21) This denition is given in 6] for the case of a continuously dierentiable Q1 and a real scalar product. If Q1 2C1, we may make use of

Q1(t)y =Q1(t)V(t)y =Q1(t)0(t)x=;Q01(t)(t)x=;Q01(t)PP1(t)x and arrive at the same expression as used in 6] instead of (2.21).

By decoupling A(t)y+B(t)x= 0, we nd that (2.21) leads to y+M(t)(t)x= 0 V(t)x= 0 Px= (t)x:

Therefore, (2.20) reads then

h;M(t)(t)x (t)xi;j(t)xj2

which means a contractivity condition for the inherent regular ODE relative to the invariant subspace im (t) =PS1(t).

19

(20)

3 Analyzing numerical integration methods for linear index-2 DAEs

In this section we derive conditions for preserving stability properties of numerical methods when solving index-2 linear DAEs. Let the conditions of Theorem 2.2 be fullled in the following. As in 6] we assume N(t) to be constant, otherwise it is well known that those methods will fail. We shall consider the discrete version of the decoupling of the last section for two of the most important families of methods.

An analogous analysis was made in 6] using the standard approach mentioned in Section 2.1. In that paper they obtained as a sucient condition for preserving the stability properties that both subspaces N1 and S1 have to be time invariant. By the discrete decoupling with the new approach described in Section 2.2, we shall get a weaker condition as in 6]. Namely, the time invariance of S1 will do.

3.1 BDF

Consider the homogeneous system

A(t)x0(t) +B(t)x(t) = 0 t2J := t0 1): (3.1) For homogeneous index-2 tractable DAEs (3.1), the solutions are given by the ex- pression

x= (I;QP1G;12 BPP1;QQ1V0PP1)PP1u= canu (3.2) where u solves the regular ODE

u0 ;0u;PP1V0u+PP1G;12 BPP1u= 0 (3.3) with u(t0)2Im(t0). Suppose P0 = 0.

If the projector V is time invariant, or equivalently, = P(I;V) is so, this solution representation simplies to

x= (I;QP1G;12 BPP1)PP1u (3.4)

u0+PP1G;12 BPP1u= 0: (3.5)

Given the step size h >0,ti =t0+ih, i2IN. The BDF applied to (3.1) reads AiXk

j=0jxi;j +hBixi = 0 ik (3.6)

where the starting values x0 ::: xk;1 are supposed to be known.

First, we multiply by ((WB)+W)i and obtain

Vixi = 0 ik (3.7)

20

Referenzen

ÄHNLICHE DOKUMENTE

Based on OGCM circulations achieved under restoring times of 30 days and 150 days we analyzed the pro- cesses which lead to intermittent convection and to the sensitivity of deep

10 were combined, the category 6 being omitted. I f this category is included, dichotomizing the response scale, 26 per cent support is obtained, a figure approaching the 29 per

We determine the phase stability of bridgmanite and LiNbO 3 ‐ type phase as a function of FeAlO 3 content and the maximum solubility of the FeAlO 3 component in bridgmanite..

To stave off collapse or violent regime change, Algeria needs deep political and eco- nomic reforms conducive to sustainable and equitable economic expansion, increased

Colloid stability/flocculation/coagulation/ are controlled by the relative magnitudes of the van der Waals and the Coulombic forces. aqueous As 2 S 3 sol with increasing

Also, the problem of determining the minimum number of mutually non overlapping con- gruent copies of a given disk which can form a limited snake is very complicated.. The only

The comparison of structural properties determined by X-ray analysis within a se- ries of related complexes shows that the Ru–N bond lengths to the coordinated bipyridines are

The reactions between ruthenium complexes containing different biimidazole-type ligands with the sulfate dianion, however, show a strong correlation be- tween the substituents at