Approximate inertial manifolds for the Kuramoto-Sivashinsky equation: Analysis and computations

Approximate inertial manifolds for the Kuramoto-Sivashinsky equation: Analysis and computations

Physica D 44 (1990) 38-60 North-Holland APPROXIMATE INERTIAL MANIFOLDS FOR THE KURAMOTO-SIVASHINSKY EQUATION: ANALYSIS AND COMPUTATIONS M.S. J O L L ...

1MB Sizes 0 Downloads 18 Views

Physica D 44 (1990) 38-60 North-Holland

APPROXIMATE INERTIAL MANIFOLDS FOR THE KURAMOTO-SIVASHINSKY EQUATION: ANALYSIS AND COMPUTATIONS M.S. J O L L Y a'l, I.G. K E V R E K I D I S a'b and E.S. T I T I c'2 aprogram in Applied and Computational Mathematics, Princeton University, Princeton, NJ 08544, USA bDepartment of Chemical Engineering, Princeton University, Princeton, NJ 08544, USA CMathematical Sciences Institute, Cornell University, Ithaca, NY 14853, USA Received 12 May 1989 Revised manuscript received 23 January 1990 Communicated by J. Guckenheimer

We evaluate several alternative methods for the approximation of inertial manifolds for the one-dimensional Kuramoto-Sivashinskyequation (KSE). A method motivated by the dynamicsand originallydeveloped for the Navier-Stokes equation is adapted for the KSE. Rigorous error estimates are obtained and compared to those of other methods introduced in the literature. Formal relationships between these other methods and the one introduced here are established. Numerical bifurcation diagrams of the various approximate inertial forms for the KSE are presented. We discuss the correspondence between the rigorous error estimates and the accuracy of the computational results. These methods can be adapted to other dissipative partial differential equations.

1. Introduction The traditional approach to the computational study of the dynamic behavior of partial differential equations (PDEs) relies on their approximat i o n - t h r o u g h discretization techniques, such as finite difference, finite element or spectral metho d s - b y finite dimensional dynamical systems. While an increasingly fine discretization is required to capture more accurately the essential features of the dynamics, it is known that for certain evolution P D E s the long-time behavior is relatively low-dimensional (see e.g. refs. [28, 29, 56] and the references therein). This is the case for dissipative P D E s whose trajectories enter and eventually remain in an absorbing ball contained 1Currently at the Institute for Mathematics and its Applications, Universityof Minnesota, Minneapolis, MN 55455, USA; on leave from Indiana University, Department of Mathematics, Bloomington, IN 47405, USA. 2Current address: Department of Mathematics, University of California, Irvine, CA 92717, USA. 0167-2789/90/$03.50 © Elsevier Science Publishers B.V. (North-Holland)

in the appropriate phase space. The evolution of this absorbing ball as a set, leads to the notion of the global attractor ~¢, strictly defined as the to-limit set of the ball, which under additional assumptions is nonempty, compact, and invariant [3]. While it is known in certain cases that the set d has a finite Hausdorff dimension (see e.g. refs. [8, 12, 22, 38, 40]), it may be quite complicated topologically and attract solutions very slowly. The theory of inertial manifolds allows us to reduce the long-time behavior of the P D E to that of a finite-dimensional dynamical system. An inertial manifold is defined as a finite-dimensional, positively invariant Lipschitz manifold which exponentially attracts all trajectories, and thus contains the global attractor. The flow restricted to such a manifold is equivalent to that of a finite system of ordinary differential equations called an inertial form. The existence of inertial manifolds has been established for a growing list of dissipative PDEs modeling physical systems, including the C a h n -

M.S. Jolly et al. / Inertial manifolds for the Kuramoto-Sivashinsky equation

Hilliard (phase transitions) [10, 49], GinzburgLandau (hydrodynamic instabilities) [7, 14, 24], Swift-Hohenberg (convection) [52], several reaction-diffusion equations [10, 32, 42, 39], damped Hamiltonian systems [46], along with the Kuramoto-Sivashinsky equation (KSE) (interfacial instabilities, wrinkled flame fronts) [10, 11, 18, 19]. It is still an open question whether the very equation which recently motivated the theory, the Navier-Stokes, has an inertial manifold. Even when existence of an inertial manifold can be established, the theory does not provide it in an explicit form. Thus, in order to implement an inertial form computationally, an approximation is necessary. A number of approximation schemes have been introduced in the literature. The approximating manifolds can be grouped into two categories: those approximating a true inertial manifold [15, 20, 41, 42], a n d those approximating the global attractor [17, 23, 26, 55, 57-59]. In either case we refer to the manifold as an approximate inertial manifold (AIM). These latter AIMs have been used independent of the existence of an inertial manifold, for example for the Navier-Stokes equation. This is the approach taken in this paper, even though the scheme is developed for a system which has an inertial manifold. Even though one would prefer to approximate an inertial manifold of the lowest possible dimension, it is possible that there are other welldefined inertial manifolds of higher dimension. In fact, to decrease the approximation errors, one often considers approximating these higherdimensional inertial manifolds. Taking the approximating system as a perturbation, Sell [51] has recently shown that hyperbolic sets are preserved when the perturbation is small enough. In view of the above, by using progressively more modes, and thus reducing the approximation error, we expect the essential dynamic behavior of the finite-dimensional systems generated by these schemes to converge to that of the PDE. The main advantage over traditional discretizations would be an expected higher rate of convergence.

39

In this paper we adapt the method developed in refs. [58, 59] for the Navier-Stokes equation to one for the KSE. The KSE has recently attracted a considerable amount of analytic and computational interest, including the papers of Amick and McLeod [1], Armbruster et al. [2], Chang [5], Chen and Chang [6], Green and Kim [27], Hyman and Nicolaenko [30], Hyman et al. [31], Foias et al. [16], Kevrekidis etal. [35], Michelson [45], Nicolaenko and Scheurer [47], Nicolaenko et al. [48], Scovel et al. [50], Troy [60]. Not only is the KSE known to have an inertial manifold, but the numerical evidence strongly suggests that this manifold is three dimensional over a parameter range that includes a great variety of bifurcations, a claim which is yet to be proved rigorously. Thus we chose the KSE as an illustrative example for this study both for its physical interest and the richness of its low-dimensional dynamics. The paper is divided into two main parts, analytic and computational. The first part is devoted to the analysis of an AIM passing through all the steady states of the KSE. After expressing the KSE in functional form below, we derive some preliminary results regarding its solutions in section 2. More specifically, we obtain precise estimates on a region of analyticity in time for the complexified equation. Using the Cauchy formula we then obtain global bounds on the time derivative of the exact solution on the region of analyticity, which contains the positive real axis (positive time). These bounds play a very essential role in the error analysis that is presented in section 3. In section 3 we introduce an AIM which passes through all the stationary solutions, and show in theorem 3.3 that all trajectories of the KSE approach a thin layer about this AIM at an exponential rate. In section 4 we discuss the motivation behind some other methods of approximation in the literature, their formal relation with various stages of the method presented in section 3, and tabulate the associated errors. Section 5 comprises the computational part of the work. There we compare bifurcation diagrams obtained using various three-dimensional

40

M.S. Jolly et al. /Inertial manifolds for the Kuramoto-Sivashinsky equation

AIMs to each other and to those obtained using traditional Galerkin approximations of increasingly higher dimension up to a dimension where the diagrams appear to have converged. We present complete steady state bifurcation diagrams and tabulate all steady state bifurcations. Diagrams including certain periodic solutions are also presented. We find that, even for such lowdimensional approximations, there is good correspondence between the rigorous error estimates and the qualitative accuracy of the diagrams. We conclude the interpretation of the computational results with the discussion of a particular side effect the approximations appear to have on the size of the absorbing ball, We treat the causes of this side effect in detail in a subsequent work where we also suggest a remedy for the problem [33].

1.1. Functional setting and notation We consider the renormalized KuramotoSivashinsky (KS) equation ~U ~4U ~2U ~U a"t- -~- OX4 q- aX 2 -~ U'~-'~

O,

( x , t ) e R X R +, (KS)

u(x,O)=uo(x), xe~, subject to periodic boundary conditions

u(x,t)=u(x+L,t),

L>0.

Let Hper((0, L)) denote the subspace of the Sobolev space Hm((0, L)) consisting of functions which, along with all their derivatives up to order m - 1, are periodic with period L. It is known that for every u 0 e L2((0, L)) which is periodic with period L, there exists a unique solution u(x, t) e H4er((0, L)) for all t > 0 (see refs. [47, 53]). Moreover, if we assume in addition that u 0 is an odd function (i.e. Uo(X)= - u o ( L - x ) , a.e. in R), then the corresponding solution u(x, t) is also an odd function (u(x, t)= - u ( L - x , t ) , a.e. 'in R).

For (KS) restricted to the invariant subspace of odd functions, it has been shown by Nicolaenko et al. [48] (see also ref. [18]) that for every p > 0 there exists a time T*(p) > 0 such that whenever IlUoIIL= _< p, then Ilu(t) IlL2 <_Po - CoL5/2,

(1.1)

IlVu(t) IlL2 < Pl = c1L7/2,

(1.2)

for all t > T*. The constants c o and c I (as well as c 2, c 3. . . . to follow) are universal in that they are independent of the problem parameter L and the initial value u 0. To date rigorous studies of the long-time dynamics of the Kuramoto-Sivashinsky equation have been limited to the invariant subspace of odd functions as the inequalities (1.1), (1.2) are essential in showing the existence of a compact global attractor and subsequently in estimating the dimensions of the global attractor and inertial manifolds (see e.g. refs. [48, 18, 10, 11]). Should (1.1) and (1.2) be rigorously extended to more general initial values, then the results of these studies could enjoy similar extensions. We should mention though that the only rigorous result in this direction indicates that for general initial data the solution would never grow faster than exponential in time (see ref. [53]). On the other hand, all the numerical simulations that have been done so far indic~ite that the solutions are bounded. With this in mind, in this work we

restrict ourselves to the odd case. The Kuramoto-Sivashinsky equation is equivalent to the functional differential equation (see e.g. refs. [18, 56]) du

d-7+Au-A1/2u+B(u,u)=O,

uGH,

where A = a4/ax 4 with domain D ( A ) = H4er((0, L ) ) , B is the bilinear operator defined by av

B(u,v) =u-~,

forall u , v e H l p e , ( ( O , L ) ) ,

(1.3)

M.S. Jolly et al. /Inertial manifolds for the Kuramoto-Sivashinsky equation

and

by parts, that

H = {u ~ L 2 ( ( O , L ) ) I u ( x , t ) = u ( x + L , t ) ,

(B(u,u),u)=O

u(x,t) = -u(L -x,t),

=

~

L)). (1.5)

x ~ R}.

Within H we will denote the L 2 inner product by ( . , . ) and corresponding norm by I" I. Since it will occur frequently we will also define

Ilull ~ IZl/4u]

forallu~Hlr((0,

41



Since A is a self-adjoint positive operator whose inverse A - ' : L2((0, L ) ) ~ L2((0, L ) ) is compact (by Rellich's lemma) it follows that the space H has a complete orthonormal basis consisting of eigenfunctions of A, {wj~= 1, where wy(x) = s i n ( 2 ~ j x / L ) corresponding to eigenvalues ,lj = (2,trilL) 4 for j = 1, 2 . . . . . Let P =Pm be the orthogonal projection from H onto span{wl, w z .... , wm} and Q = Qm = I - P. Since P and Q commute with A and its powers, we may split eq. (1.1) as

2. Preliminary results Applying the techniques developed by Foias and T6mam [22] for the Navier-Stokes equations, we show in the first paragraph that the solution to (1.3) is analytic in time with values in D(A), the domain of A. In the second paragraph we follow the approach of Foias et al. [17] for the Navier-Stokes equations, and combine the analyticity in time with the Cauchy formula to derive several key estimates for :the time derivative. We will apply these estimates here and in subsequent works on the KSE. To be more precise, the estimate (2.13) of theorem 2.2 below plays an essential role in measuring the uncertainty of predicting the location of the global attractor; as is illustrated in theorem 3.3. We should also remark that this property, analyticity in time of solutions, holds for a more general class of dissipative systems. Recently, Foias and T6mam [23] have used this property to construct algebraic sets approximating the global attractors of these systems.

dp + A p - A 1 / 2 p + P B ( u , u) = O, p - - P u , dt (1.3 9 ) dq d'--7 + Aq - a l / 2 q + Q n ( u, u ) = 0,

q -----a u .

(1.3q) Let us recall Agrnon's inequality in one space dimension,

2.1. Analyticity in time For this purpose we consider the complexified equation du d---~+ Au - A l / 2 u + B ( u , u) = O, u~H,

~C,

withRe~>0,

(2.1)

Ilull~ ~ V~-lull/211ulll/2, which implies

I( n ( u , o ) , w ) l

~ v/-21ull/211ulll/21lolllwl.

(1.4)

And in addition one can verify, using integration

where H, D(A), A, and B now denote the complexified extensions of the related spaces and operators. While it is easy to see that (1.4) still holds for the complexified case, the same cannot be said for (1.5). We must actually estimate (B(u, u), u) in the proof below.

42

M.S. Jolly et aL /Inertial manifolds for the Kuramoto-Sivashinsky equation

Theorem 2.1. Let p > 0 and J be a region in the complex plane defined as the union of all the open disks, with radius 3, centered at 6 + ~- for all ~"> 0, where

(by interpolation) × x/2-A~-x/Slu121A1/2u{. By Young's inequality we have

I(n(u,u),u)l 8p + 2A~/4"

lul 4

Suppose u(~) is a solution to (2.1) on a domain containing the positive real axis, such that lu(t)l
forallt>0.

(2.2)

Then the solution u(~') exists in ~ / and the mapping u: J ~ D ( A ) is analytic and satisfies

lu(~)l _<2p f o r a l l ~ E ~ a, ~ t t)

< _~

for all t > ~1 ,

(2.3) (2.4)

+ i COS0 [al/2ul 2.

--< A1/4 COS0

(2.8)

It follows from (2.6), (2.7) and (2.8) that __

2

d lul2< l u l 4 + c o s 0 lul 2 ds - •11/4 cos 0

Setting y ( s ) = lu(s ei°)12, Yo = luol 2 we have dy

d s -< a y ( y + t3), where

IAu(t)[ < K I ( p ) -- ~32p + 4p + 10p 13/5,

2 1 for all t > ~6.

(2.5)

=

cos o '

= l x Y ' cos

Proof. Following the approach in ref. [22] (see

Direct integration leads one to

also ref. [17]), we take the inner product of (2.1) with u, multiply the result by exp(i0) and take the real part to obtain

- - Y < s c o s O - - Yo y+/3 yo+/3"

o.

Seeking an upper bound on s to ensure that

lU(S e io) 12

+ 21( B ( u , u ) , u ) l .

(2.6)

By means of interpolation and Young's inequality we obtain

IA1/4u12< lul IA1/Eul

(2.9)

y < 2y 0,

_< 2cos 0 ( - Ial/2ul2 + Ihl/4ul 2)

we note that since the function z(z+fl) -1 is increasing on z ~ (0, oo), relation (2.9) will hold provided Y0 2Yo s cos Oy-~--~ _< 2Yo+ fl

_ I i ~2+ 1 --1/2 2 --<~IU[ ~12't Ul •

(2.7)

or equivalently

Applying (1.4) we find that

1

s<

log

2y o + 13

I( n(u, u), u) I ~ v~-)u la/211ul13/2 x/2A11/alul Ilull2

t3

"-- ~ - - ~ l o g ( 1 + 2yo + fl )-

M.S. Jolly et al. /Inertial manifolds for the Kuramoto-Sivashinsky equation

Using the fact that

Combining the interpolation

1 log(1 + z ) > ~z,

llull~ [ul3/4lAu[I/4

z~ (0,1) ,

we find that

43

with Young's inequalitywe obtainfrom (2.10)

t~ co--o-~l°g ( 1 + 2yo +/3 )

llZul _< d--d-~t + alul + v~lul13/SlAul 3/8.

}kl/4

_> COS0

8Y0 + 2A~/4 COS20

We use Y0ung's inequality again to obtain

_ vc2-6 cos O.

½lZul <_

d_~t t +

l l u l + g5 M 24/51U[ 13/5 + -ff 3 IAu l .

Thus, within the lemniscate _E= {g"= s ei°ls = re2-6cos O, - Ir/2 < 0 < ~r/2} relation (2.9) holds. It follows from the methods developed by Foias and T6mam in ref. [22] that the solution to (2.1)exists and is analytic within _E. By assumption (2.2) the above argument can be applied to the solution of (2.1) with initial value taken as u(t o) at any t o > 0 to conclude that the solution is analytic and satisfies (2.9) within the translated lemniscate .E+ t 0. One can easily check that the union of all such lemniscates for t o > 0 contains ~/, proving (2.3). To prove (2.4) we apply the Cauchy integral formula to the circle F with radius 16 centered at 1 Observing (2.2) any point on the real axis t > 38. we obtain

d_d_~ t =

~

1 fr ( u(~') d~" < t-~')

4p "~-"

To prove (2.5) note that as a direct consequence of (1.3) we have

Ihul ~ -~tl + Ial:ul + In(u,u)l. From (1.4) and the Cauchy-Schwarz inequality we have

2.2. Behavior of higher modes In this section we show that every orbit for the Kuramoto-Sivashinsky equation eventually enters a thin neighborhood of the linear manifold I'mH. This implies that in particular the global attractor lies within this neighborhood. In the next theorem we estimate the thickness of this neighborhood and the rate of its exponential attraction. The interpretation of the estimate (2.12) is that the energy contained in the modes with high wave numbers, and accordingly in the small scale structures, is exponentially dissipated and asymptotically very small, The estimate in (2.13), on the other hand, implies that these small scale structures are asymptotically slow with respect to the large scales, even though they may appear fast with respect to their own small scales. In other words, while Idq/dtl may be large when compared to [q I, it is small when compared to Ipl.

Theorem 2.2. Let m be large enough so that Am+ 1 > max(16,8(2PoP~)l/2h;1/'),

[ull/2lal/2u[1/2 Iaul < ud._d7 +

+ ~/~ lull/2llull 3/2.

Applying assumption (2.2) and (2.4), we conclude (2.5).

(2.10)

(2.11)

where P0, Pl are as in (1.1), (1.2). Then for every p > 0, every solution u(t) = p(t) + q(t) of (1.3)

44

M.S. Jolly et al. / Inertial manifolds for the Kuramoto-Sivashinsky equation

-i If we set p = (8p0Pl)3 i/2 Am+i, then from (2.12) we find that

with lu(O) l ~ p satisfies

Iq(t) [2 < [q(T*)[ z e x p [ - ½Am+i(t - T*)] Iq(t)[ < p

3 -2

+ 4poPiAm+ 1, for all t > T * ( p ) ,

(2.12)

forall t > T**(p).

An argument similar to that used for (2.4) can be used here to show that

where T* is as for (1.1), (1.2). Moreover, ~ t t) ~tt)

__ T * * ( p ) ,

(2.13)

for a l l t > T**,

(2.15)

where ~(p) is as in theorem 2.1. Since from (2.11) we have 8(p) > ¼, (2.13) follows.

where

r**(p) = max{T*(p) + A~,l+lllog(4p~/A2+lPo)l, ~},

and g 2 = 32(2pop3) 1/2. Proof. Let t > T*(p). Taking the inner product of (1.3q)with q we have 1 d 2 dt Iql2 + Ihl/2q12

<_ Ial/4ql2 +

< / ~ (4p p)

Since on the global attractor we have for all t ~ R that solutions exist and (1.1), (1.2) hold, the following is an immediate consequence of theorem 2.2. Corollary 2.3. Let m be large enough to satisfy (2.11). Then for every solution u ( t ) = p ( t ) + q(t) on the global attractor we have [q(t) 12-< -4p°p3 '5---

I( B(u,u),q)l

for a l l t ~ R

Am+l

<- Aml/121Ai/2ql 2 + I ( B ( u , u), q ) l .

(2.14)

and

Using (1.1), (1.2), and (1.4) we find that ~tt)

_
f o r all t ~ R.

I ( n ( u , u), q) I -< v~-Iull/211ull3/21ql ,/~-~1/2,~3/2~ - i/z i Ai/2~, I <-~ v ~ Y O

P'i

"*m+l i~*

"11

3. Steady approximate inertial manifold

(by Young's inequality) --1+ i + -~lA1/2q[ 1 -~


2.

Now (2.14) becomes < 2Pop3Am 1+i" -dtd - iql 2 + 9(1 ~ \ ~ - - ' *~t m-i/2~lai/2,~l + i ] Ixi ' . i l 2 -Using (2.11) we get d 1 3 -1 1. -dt- Iql2 + ~Am+llql 2 _< 2poPlAm+ By the standard Gronwall inequality we obtain (2.12).

We adapt the method used for the NavierStokes equations by Titi [58, 59] to introduce a similar approximate inertial manifold for the Kuramoto-Sivashinsky equation. Let

= { p ~ P H : Ilpll ~ 2p1}, ~.L = [q ~ QH: Ilqll ~ 2 p l ] . As for the Navier-Stokes equations we follow Foias and T~mam [21] and Titi [58, 59] to show that there exists a mapping ~s: ~ __+Q H satisfy-

M.S. Jolly et al. / Inertial manifolds for the Kuramoto-Sivashinsky equation

ing

tiating (3.3) one obtains

OT,,(q),7 a------q--

+ ¢ ' ( v ) ) = o,

+Qn(p + p~.

Theorem 3.1. l e t m be large enough so that Am +1 -> max( 4r~, rl2/4p 12),

(3.2)

where +

~J~-r~2,~ -1/8,~"*rn-1/4 ~ v ~ P'I'*I + 1

,

r 2 = 1 + 8PlAi-1/SAml+/14, with P0, Pl as in (1.1), (1.2). Then there exists a unique mapping ¢~: ~ - - , Q H satisfying (3.1), whose graph defines a C-analytic manifold. Moreover,

IlCS(p)ll~A~+lr 1, for all p ~ ~ .

- A - 1 Q [ B( 77' P + q)

+B(p +q,n)].

(3.1)

The graph of Cs, denoted ..~/s,contains all the steady states of (1.3). For this reason we will call the manifold .~s the steady approximate inertial manifold.

r I = 2Pl

45

Using again (1.4) we have that

aq

_ Am1/1211mll÷ 8P1A~-1/SAm3+/141bTII A,L/12r211II ½16711.

It follows that Tv is a contraction mapping. Thus for any p ~ ~ there exists a unique fixed point of Tp we denote by ¢~(p), which defines ¢~: ~---> ± satisfying (3.1). As in ref. [21] it follows that ¢ s is a C-analytic manifold. The mapping ¢s, which is given by the implicit relation (3.1), is our candidate for an AIM. In the next theorem we introduce a sequence of simple explicit approximating functions for Cs. These approximating functions will also be associated with AIMs, which are the ones used in part of our actual computations in section 5.

Theorem 3.2. l e t m be large enough so that (3.2) holds. Define

Proof. For each p ~ ~ we define ¢0(p)=0

Tp(q) -..~A-1/2q - A - I Q B ( p

for a l l p ~ ,

+ q , , + q),

for all q ~ ~ _L

and (3.3) ¢i+1(P) = Tp(¢i(P))

We begin by showing that T,: ~ ± --* ~ 1. From (3.3) we have

for all p ~ ~i~, and for i = 1,2,3 . . . . . Then

IITp(q) II -< IA -

~/4ql

+

~

- 3 / 4 Q B ( p + q, p + q ) I.

I1¢111---~ " Ic ~-3/4 J-x 3,'~.m + 1

(3.4)

,

Using (1.4), it follows that

and

IITp(q)ll < 2plAto 1+/12+ 8~/2-plzA1 l/8~t'~m+l-3/4

lieS(p) - ¢ , ( P ) II ~ 2K3a'A~,3+/14,

= A,:1+/12rl < 2Pl. We now show that Tp is a contraction. Differen-

where r 3

=

4~-li-'/Sp~,

a

=

r2Xml/12.

(3.5)

M.S. Jolly et al. /Inertial manifolds for the Kuramoto-Sivashinsky equation

46

Proof. Let p ~ ~ be fixed. Using (1.4) we find

and

that ][q(t) - ~i(p(t))ll <--2K2Am+ -7/4x + 2K3a ih ,-3/4 ,+l,

[l~l(p)l[- [A3/4QB(p,P)I

for a l l t > T * *

and

i=1,2 .....

(3.8)

<_A~3+/141B(p, p) I where T** and K 2 are as in theorem 2.2, and K 3 and ot are as in theorem 3.2.

_<4~-A ~-X/Sp12/~m3+/14 proving (3.4). Since Tp is a contraction mapping, with contraction constant a (see proof of theorem 3.1) we have

lu(t)l-
and

Ilu(/)ll-
for all t > T**(p).

II~k+ X(P) -- ~ k ( P ) II ~ akll~x(P) -- 4 0 ( P ) II = akll~l(p)l[

Proof. Because of (1.1) and (1.2) we have

In particular qbS(p(t)) is well defined and satisfies 114,S(p(t))ll_< 2pl (see theorem 3.1). Let A ( t ) = q ( t ) - ~ S ( p ( t ) ) , for t > T**. From (1.3q) and (3.1) we have

for k = 0,1,2 . . . . .

Since oc

qbS(P)

-- Cibi(P) =

E

[qbk+l(P)

-- ¢ J ~ k ( P ) ] ,

AA -Ax/2a + Q[ B( A , p + ~S(p) ) + B(u,A)]

k=i

dq

we have

+-d-T = 0 . co

II~S(p)

-

Using (1.4) we get

~i(p)ll-< E akll~x(P)ll k=i

IAAI < IAX/2al

2a'll~,(p)ll.

+ v~-Ia 11/211al11/211p + ~S(p)II We now substitute (3.4) to get (3.5).

+ Following Titi [58, 59] we show in the next theorem that all trajectories of (1.3) are attracted exponentially fast to a layer about graph ( ~ 0 , which is significantly thinner than that about PmH given in theorem 2.2.

Theorem 3.3. Let m be large enough so that (2.11), (3.2) and A--X/2m+I"+ A,~ ~t--7/8 -[- J~-,~l/2,~l/2Jt - 3 / 4 < 1 -r~X"m+X v~YO /"X " * m + l --

"~1('~'~

i

ll,211u:ll

ll +

From the above and (2.13) we obtain IAAI -< (~-1/2 v-,,+x + 4,°xX,7,,7/~8 + ,/~-~1/2,,~1/21 - 3/4 ~] IA zi I + K 2 A m l+ xv'~'P'0 V'l "~m+l

We now substitute (3.6) to reach (3.7). The estimate (3.8) follows directly from (3.7) and (3.5).

hold. Then for every p > 0 and for every solution u(t) = p ( t ) + q(t) of (1.3) with lu(0)l _
From estimates (3.7) and (3.8) we infer that the graph of qb2 is as close to the global attractor as the steady AIM, graph(qb0. We will refer to the manifold given by the graph of qO2 as the

[A(q(t) - ~ ( p ( t ) ) )

pseudo-steady approximate Menial manifold.

I < 2g2Arn+X ,-1

(3.7)

M.S. Jolly et al. / Inertial manifolds for the Kuramoto -Sivashinsky equation

Remark 3.4. The formal derivation of Tp from the implicit relation

Aq-A1/Zq + QB(p + q,p + q) = 0 amounts to simply bringing both the destabilizing linear t e r m al/2q, along with the bilinear term, to the right-hand side before operating on both sides by A-1. An alternative approach would be to bring over only the bilinear term, and then operate by ((A-AX/2)Q) -1, leading to the mapping

Tp( q) = - ( ( A - A ! / 2 ) Q ) - I Q B ( p + q, p + q). We remark that, for large enough m, one can show that Tp is also a contraction mapping defined on ,fi~.t. Accordingly, one can approximate ~s by the successive iterates of Tp, namely ^

4. Analytic comparisons of approximate inertial manifolds In this section we compare both the motivation and error estimates for three AIMs in the literature with three new ones introduced here as variants on the pseudo-steady AIM. For the purpose of this discussion it is convenient to consider the general evolution equation du

d----[+ Au + F(u) -- 0.

o(p) -= o, f o r p ~ ~ , and i = 1,2 . . . . . In addition, one can restate the analogue of theorems 3.2, 3.3, with ~i replacing ~,.. In doing this, however, there is no gain in the rate of approximation of ~i over that of ~;. Moreover, ((A-A1/2)Qm)-I cannot be inverted for the parameter L -- 2.rr(m + 1), which is precisely the value at which the trivial solution undergoes a pitchfork bifurcation making the dimension of the unstable manifold of the origin m + 1 dimensional. While beyond this parameter value there could be no true inertial manifold of dimension m, we will see in section 5 one instance in which the m-dimensional version of ~2 continues to capture important qualitative behavior. Nevertheless, ~1 does have the advantage of involving a major instability term which is not present in ~1. We will not address this issue in detail in this work, but we will comment on it later in section 4.1.

(4.1)

In the case of the KSE we lump -AX/2u and B(u, u) in (1.3) together as the nonlinear term F(u). We will often refer to the approximate inertial form (AIF)

dp + a p + e F ( P + ~ a ( p ) ) = O , dt

p~eH, (4.2)

^

=

47

where the AIM is the graph of ~a: P H ~ QH. The error in all the methods is described in terms of the maximum vertical distance in the H-norm between the global attractor ~¢ and the graph(~a). More precisely, for u - - p + q a~', we have Error(~a) = max Iq~a(p) - q l.

^

^

Since all the error estimates in this paper are of the form kAm~+X for some positive constants k and y, both independent of m, we will simply speak of the order of the error as am~+l.

4.1. The traditional Galerkin AIM The simplest of all AIMs is a flat one. In theorem 2.2 we have seen that for m large enough there is a layer about PmH, with thickness of -1 order Am+l, which attracts every trajectory exponentially. To employ this linear manifold PmH, amounts to simply setting ~ a - 0 in (4.2), and gives the traditional Galerkin approximation of (4.1) associated with the eigenfunctions of the

M.S. Jolly et al. / Inertial manifolds for the Kuramoto-Sivashinsky equation

48

operator A. We use this as the standard to compare against any other AIF. More sophisticated AIMs have some curvature due to nonlinear terms in q~a. For this reason the following AIFs are often referred to as "nonlinear" Galerkin methods of approximating (4.1).

4.2. An earlier approach to

(1) 1

The following method was introduced by Foias, Manley and T~mam for the 2D Navier-Stokes equations in bounded domains. Projecting onto the space of divergence free functions, one can show that the Navier-Stokes equations with no slip or with periodic boundary conditions reduce to a special evolution equation (cf. refs. [10, 37, 54]), du

d--'-t+Au + B ( u , u ) =f.

(4.3)

Splitting (4.3) as was done for (1.3), and expanding the bilinear term, we have for the higher modes dq

4.3. The Euler-Galerkin AIM We now turn to a method which approximates a true inertial manifold, rather than just the global attractor. The key to the Hadamard approach to prove existence of inertial manifolds is the evolution forward in time starting with appropriate initial data [10, 11, 39]. For example, the semi-flow induced by the solution of (4.1) takes an initial manifold ~/0, under appropriate assumptions, to a true inertial manifold ~ as t ~ oo. Since .Z/is exponentially attracting, after a relatively short time interval ~', . ~ , the image under the semi-flow of . d 0, can be shown to be already very close to .~. This is the motivation behind the AIM in ref. [20]. The key in that approach is to approximate the true evolution with implicit Euler integration for small time r. This yields, for ~" small enough, an AIM q = ~ ( p ) satisfying

q = -~'(~'A + I ) - ' Q F ( p +q), which is the fixed point of the contraction mapping

+Aq + Q [ B ( p , p ) + B ( q , p ) + B ( p , q ) ~ff-: q~, - ~ ' ( I + r A Q ) - I Q F ( p + q ) .

+ B ( q , q)] = Qf. It is shown in ref. [17] that for large enough m and uffip+q on z~' one has that Idq/dtl, IQB(q,p)I, IQn(p,q)l, and Ian(q,q)l are relatively small. Neglecting these terms, they obtain the AIM given simply by the graph of

(Pl( P) = - A - 1 Q [ f _ B( p, p ) ] ,

(4.4)

which has an error of order A~3+/12. Here, however, the eigenvalues of A, the Stokes operator, only grow linearly [44]. Notice the similarity between qfil in (4.4) and ~1 in theorem 3.2. There we also estimated an error on the order of Am-3/2 +l , except that for the KSE the eigenvalues grow like m 4 (see section 1).

(4.5)

The first iterate of the mapping 2V- with initial guess .~o=Pm H, i.e. q = 0, yields an explicit function qtl,~: PmH ~ QmH given by

qtl,~(p ) = - r ( I + ~ ' A Q ) - X Q F ( p ) .

(4.6)

It is shown in ref. [20] that the maximal vertical distance between graph(qtl,~) and graph(g'~) is comparable to that between g r a p h ( ~ ) and .£~, which is, for appropriately small r, comparable to that between d(~ and .~. In the case of the KSE this error is of order A~7/14. We refer to the manifold defined by (4.6) as the Euler AIM and the AIF (4.2) with qb, __ ~1,, as the Euler-Galerk/n method.

M.S. Jolly et al. / Inertial manifolds for the Kuramoto-Sivashimky equation

49

Observe that as z---, % so that the Euler integration is no longer rigorously justified, we have that ~ --->~s and ~1,~ ~ ~1, where ~1 is as in theorem 3.2.

For (ii) we start by observing that again by (3.4)

4.4. New methods: variations of the pseudo-steady AIM

Estimating Ilpxll~, we have by Agmon's inequality

<~C3p~/2 Ipl 1/4 IApl 1/4

= A-1/2CI)l( p )

< c3P~/2p 1/4 lAp[ 1/4

-A-1Q[ B ( p , p ) + B( qOl,p) +B(p,

~1) + B(~I,

~1) ] .

(4.7)

In the proposition below we estimate certain terms in q~z which are later dropped to produce three new AIMs.

Proposition 4.1. Let m be large enough to satisfy (2.11). Then for every u = p + q ~ ~ ing estimates hold:

the follow-

[A-1QmB(+I(p),~I(p))[

(i)

9 J ~ - / ( 2 ) , - 21/8 -~< ~ v ~ 3 Z t m + 1 ,

(ii)

[A-1QmB(~l(p),p)l ~.. ,.~1/2~1/4,.~ 1/41 -- 2 -~<~36,1 6,0 6,2 "*m+l,

(iii)

[A-IQmB(P,~I(p))I ,/rg-~ 1/2~ 1/2~ - 7/4/,,-~< v~6,0 6"1 "*m+1"~3,

where/9 2 is as in theorem 2.1, Po and Pl are as in (1.1) and (1.2), respectively.

Proof. By Agmon's inequality and (3.4) we have IA-1Qmn(fI)l( p),

t~l(p))[

-< ,LT,'+,ll'ar'1(p) IMl'ath(p) II _< ,LT,~+i v~-ICh(p) 11/211~,(p) II3/2 -< x,7, ~+i v'2 ~,7,,~+/~8II¢,i(p) II2 _

,I - I

,]-~~

- 1/81,-

2 A - 3/2

<'*m+lV~"~*m+laX3

_<,~m2+~llPxlL.

IIp~IL ~ c3lp~l WZlp~l~/~ < cap]/2lA1/2p[1/2

Recall from theorem 3.2, the function qD2(p)

IA-1QmB(~I(p),p)[ < Am111~(P)I IlPxll®

m+l "

Since Ihpl ~P2 by theorem 2.1, we obtain estimate (ii). Finally, we have by Agmon's inequality, (3.4), (1.1) and (1.2) that [m- 1Qmn ( p, t ~ l (

P))I

A~,~+IlIplLII~I(p)II - 1 1 V/~ IPl -~
1/211PIIt/2KaA~3+/14

- 7 / 4 J'~- ,~1/2.1/2/d" -~<'tm+l v ' 6 , 0 6,1 ~x3

proving (iii). We now consider dropping some of the terms in (4.7). Since the terms in (i) and (ii) are no - 2 1, dropping either one or larger than order Am+ both will yield an AIM with the same error as for ~s. We will refer to the AIF derived by dropping only the term in (i) as "pseudo-steady II", and that obtained by dropping the terms in both (i) and (ii) as "pseudo-steady III". Dropping all three terms estimated in proposition 4.1 yields an approximation with an error dominated by the term in (iii). Thus this last variation, "pseudosteady IV", gives an error of the same order as the Euler AIM, ~-7/4 "*m+l • In fact by gathering the B(p, p) terms we find that the method formally requires just one iterate; thus it resembles the Euler-Galerkin method in form as well. We also remark that one can derive the pseudo-steady IV manifold from ~t(P), which was introduced in remark 3.4. Indeed from remark

50

M.S. Jolly et al. / Inertial manifolds for the Kuramoto-Sivashinsky equation

Table 1 Error estimates of order A,Sn~.1 for AIMs applied to the KSE. Approximation ~(p) flat (linear)

¢~1 pseudo-steady(4,2) pseudo-steadyII pseudo-steadyIII pseudo-steadyIV Euler (~1,~,)

~,

0 -A-1QB(p, p) A- I/2~1(p) - A - IQB(p + qBl(p), p + ~ I(P))

1 3/2 2

4 2 - A - t QB( ~ l( p ), q~l( P )) cb2 - A - 1QB( ~ I( p ), p + ~ I( P )) - ( A - 3/2 + A - 1)QB(p, p ) - ' r ( I + c A Q ) - 1QB(p, p)

2 2 7/4

3.4 we have

7/4

between the m-dimensional first iterate approximations (~1, Euler, pseudo-steady IV) and P2m H .

4,,(v) = - ( A -A'/2) 'QB(v,v), 5. Computational results and therefore

¢~l(P) ~'=-

AQ) -'-k/2 B ( p , p ) .

By keeping just the first two terms in this infinite geometric series, we obtain the pseudo-steady IV manifold, with a negligible remainder. In table 1 we summarize the forms of all the approximations discussed in this section as well their error estimates. It should be emphasized that these estimates in general are not sharp. An exception is the estimate for the fiat manifold, where examples are provided in refs. [20, 58, 59]. At first glance it appears that to achieve the accuracy of an m-dimensional pseudo-steady AIM, one would have to use roughly m 2 modes in a fiat (traditional Galerkin) approximation. However, this conclusion is easily disqualified in the case of the KSE. Due to the choice of basis, and the fact that the nonlinearity is quadratic, we have that Q 4 m ~ 2 ( p ) -- 0 for all p ~ PmH (a similar observation was made in the context of a center manifold approximation in ref. [2]). While this makes the method easy to implement, as there is no truncation error, it also implies that the mdimensional pseudo-steady (II, III) AIM is actually contained in P4m H, and hence could be no closer to the global attractor than the 4mdimensional linear space. The same relation holds

We now evaluate how well the various AIFs capture the dynamics of the Kuramoto-Sivashinsky equation. Rather than examine the accuracy of individual trajectories at particular parameter values, we compare entire bifurcation diagrams of stationary and periodic solutions. To ensure that no bifurcations are missed in the continuation, we also calculate the stability of the solution branches. In this comparison we concentrate on a parameter range over which the structure of elements of the global attractor suggests the existence of an inertial manifold of dimension three. It should be emphasized that as of yet, this claim is far from being verified analytically. After rescaling time and space (KS) becomes au

o-7+

4 04u

[ i~2u

0 < x < 2rr,

0U )

=0, (KS~)

where a = L 2 / r r 2. All of our computations were performed for (KS~). The bifurcation diagrams were obtained using AUTO, a bifurcation package [13] while the phase portraits were produced with an interactive code, SCIGMA [34]. 5.1. S t e a d y states

The bifurcations of the steady states for the KSE have been examined numerically in a num-

M.S. Jolly et al. / Inertial manifoldsfor the Kuramoto-Sivashinsky equation

ber of papers, including but not limited to refs. [2, 5, 6, 27, 30, 35, 36, 45]. These studies, when restricted to the space of odd solutions, concur with the results of a twelve-mode traditional Galerkin approximation over the range a ~ [0, 70] shown in fig. la. A further motivation for using twelve modes, when eight would yield nearly the same results on the interval of interest, is that as we discussed at the end of section 4, the threedimensional pseudo-steady manifolds parametrize the fourth through twelfth modes in terms of the first three. We first introduce some terminology for the elements appearing in most of the diagrams with a brief review of the situation for twelve Galerkin modes. At a = 4 , 16,36,64, the trivial solution undergoes pitchfork bifurcations leading to solution branches referred to as unimodal, bimodal, trimodal and quadrimodal, respectively. Each solution branch is symmetric in a certain subset of modes, for which the signs can be switched. We shall distinguish the two sides of each branch as either "positive" or "negative" according to the signs of the symmetric modes near criticality. Note that the two sides of each branch will be superimposed in our bifurcation diagrams where we plot the solution norm versus parameter. These primary, non-trivial branches undergo a certain pattern of secondary bifurcations according to a predictive scaling law in ref. [50]. We follow Scovel et al.'s labeling scheme. In particular, the positive side of the bimodal branch undergoes its own pitchfork bifurcations leading to the bi-tri branch at the point R2b2, and the tri-bi branch at REb 4. At REb 3 the negative side of the bimodal branch also has a pitchfork bifurcation, leading to a symmetric giant branch of solutions which apparently grow in amplitude without bound as a increases [27]. Both the tri-bi and giant solutions play critical roles in distinguishing the performances of the various nonlinear Galerkin methods. We now compare the steady state bifurcations for the twelve-mode traditional Galerkin method with various nonlinear Galerkin methods. As in-

51

dicated in fig. lb and in ref. [16], the steady state bifurcation diagram for the Euler-Galerkin method (with 7 = 0.1) is qualitatively correct up to ot = 40. Noticeably absent however, are both the tri-bi and giant branches between a = 36 and a = 64. The same is true for the pseudo-steady II Galerkin shown in fig. lc. Only the pseudo-steady Galerkin in fig. ld successfully captures one of the branches between 36 and 64, the bi-tri, albeit at a considerably inflated value of a. While we expect the corresponding bifurcations to occur at reasonable values of a, we have also searched as high as a = 300. Henceforth, when a bifurcation is said to be missing, we mean that we have not observed it up to this value of a. We remark that we carried out similar computations using the AIM given by 41, but found the results nearly indistinguishable to those for the Euler-Galerkin manifold. We also compare the results described above with lower-dimensional traditional Galerkin approximations, to determine the minimum number of traditional modes necessary to produce the essential features of the correct diagram. In figs. le, lf, lg, and lh are the diagrams for the three-, four-, five- and six-mode traditional Galerkin methods respectively. Observe that the nth mode bifurcates vertically if the traditional Galerkin truncation is of dimension < 2n, since in that case, it is a " p u r e " mode with no nonlinear balancing, as is easily seen in the equations provided in the appendix. A quantitative comparison of the steady state bifurcation values is provided in table 2. 5.2. Periodic solutions

We now turn to the time-dependent behavior of the KSE. For comparison purposes, we start with the eight-mode traditional Galerkin approximation, which appears to share the same behavior as the true KSE, for the parameter interval of interest here. The first periodic branch we encounter while increasing a, bifurcates from the positive bimodal at HB r It initially consists of

M.S. Jollyet aL / Inertialmanifoldsfor the Kuramoto-Sivashinskyequation

52 15

15

(c)

(a) Pseudo- Steady ]I

Traditional Galerkin,12 modes

n,0 Z

3modes

o

Z

Z 0 I--

Z

3

o

0

(2 ~

70

15

O(

~ ~

15

(b) Euler-Golerkin, 3modes

70 (d)

Pseudo- Steady 3 modes

Ig Z

Z

Z

Q F-

I.D .J

£ 0

Q ___.~

OJ

70

t5

I1 -------~

70

15 (e)

Traditional Galerkin 5 modes

Traditional Galerkin 3 modes

:E nO Z "7 0 I,-

0 Z Z 0

J 0 (/)

J O

I

70

15 Traditional Galerkin :E

Traditional Galerkin 6 modes

nO z

z F2D _.1 O

Q ~

70

Q ~

70

Fig. 1. Steady state bifurcation diagrams, solution norm versus parameter a. The solution norm is

Ilull2s =

L a2, where u(x,t)= L ak(t)sin(kx)" kffil

kffil

The open squares denote steady state bifurcations, the black squares denote Hopf bifurcations. In (a) bold lines indicate where both superimposed branches are stable, while . . . . . indicates where only the "negative" side is stable.

M.S. Jolly et aL /Inertial manifolds for the Kuramoto-Sivashinsky equation

,_o_s.oy:ct

7 Trodihonol Golerkin

i

8modes

53

(o)

o g

i11/ HB2 - " -"~ : " ~ 2 " - / I

I

I

I

I

130

g 8ct~ ,

1I

Ct ~

37

,

,

30

,

I

k'

/

CZ ~

37

(b)

od.es

Z•

IE o z

~

Pseudo-Steody

I

/ I--

I

I

I

I

I

a

~o 25

i

I

I

I

~

37

I

l

[

(e)

I

,I

a

Traditional Golerkin

Trad,tiono I Galerkin 3modes

~E

I

30

I

/

II

~

37

i

5 modes

I

(g )

I

I i

rr

o

o

Z

Z

o

....

g

.............

i---

g _ _

1

~

I

I

t

"""-'--'--~ I

I

t

CZ

40 I

Traditional Galerkin

4modes I

I

I

a

I

~

I

i

f

37 /

(f)

Traditional Golerkin

5:

6 modes

( h)l

n.o z z

nO z z o I-

F-

6

o o3

_o _J

%

i

1

i

(2 -------=-

i

I

i

J

40

30

J

t

i

I

C/ -------=--

it

|

37

Fig. 2. The solution norm is Ilvll~ as in fig. 1 for steady states, and ]lvlf2r= 1 rT n~ ~J. L

~ k=l

2 ak(7")d'r,

for periodic solutions, where T is the period and n is the dimension of the approximating dynamical system. ( - - - ) steady states; ( - - ) limit cycles.

M.S. Jolly et al. / Inertial manifolds for the Kuramoto-Sivashinsky equation

54

Table 2 P a r a m e t e r v a l u e s a for s e c o n d a r y s t e a d y s t a t e bifurcations. Galerkin method

R2b 1

R2b 2

R2b 3

R2b 4

R3t 1

R3t 2

R4b I

R4q I

3-mode Euler-Galerkin 3 - m o d e p s e u d o - s t e a d y II 3-mode pseudo-steady 12-mode t r a d i t i o n a l 6-mode traditional 5-mode traditional 4-mode traditional 3-mode traditional

16.103 16.130 16.131 16.140 16.140 16.140 16.140 16.000

20.590 21.928 22.009 22.556 22.547 22.828 21.124 16.000

246.14 102.90 93.913 52.891 52.721 48.493 52.067 16.0??

63.737 63.278 61.856 62.355 16.0??

36.206 36.206 36.238 36.234 36.234 36.000 36.000 36.000

63.908 50.911 46.851 36.000 36.000 36.0??

64.559 64.000 64.000 64.000 -

64.275 64.000 64.000 64.000 -

stable limit cycles, which as sets are symmetric in the odd modes. At a = 32.85 the symmetric periodic orbit becomes unstable, giving birth to two stable limit cycles, both asymmetric by themselves as sets, yet also reflections of each other in the odd modes (and hence superimposed in fig. 2a). The symmetric branch becomes extremely unstable (one Floquet multiplier eventually greater than 5000), making its continuation increasingly difficult computationally. For this reason we do not attempt to continue this branch to its conclusion. A sequence of period doublings of the asymmetric branch follow, with stability inherited by the period doubled branch at each bifurcation. We have continued the original branch of the now unstable, asymmetric solutions well past the period doublings and through a "corkscrew" like behavior indicative of a Sil'nikov loop [26, 61]. The two superimposed periodic branches bifurcating from each side of the bi-tri branch at HB 2 undergo a period doubling, after which both the original and the doubled branch exhibit Sil'nikov corkscrews. For details of the periodic behavior see ref. [4]. The diagrams for the nonlinear Galerkin methods have all the features mentioned above. Two of them are illustrated in the three-dimensional phase space of the pseudo-steady AIF. In figs. 3a-3d we depict the initial stages of the period doubling sequence of the asymmetric limit cycle culminating in a "thick", apparently chaotic orbit. An example of a periodic solution near a Sil'nikov loop is shown in fig. 4. We note that the Euler-Galerkin AIF produces artificial turning



(b)

(c)

Fig. 3. P e r i o d d o u b l i n g s e q u e n c e of a n a s y m m e t r i c limit cycle at (a) a = 32.975, (b) a = 32.989, (c) a = 32.993, a n d (d) a = 32.997, o b t a i n e d via t h r e e - m o d e p s e u d o - s t e a d y approximation.

M.S. Jolly et aL /Inertial manifolds for the Kuramoto-Sivashinsky equation

55

whose detection does not improve monotonically in the number of modes.

5.3. Artificial inflation of the absorbing ball

Fig. 4. An unstable limit cycle near a Sil'nikov loop at a 32.989 and the point in phase space (0.9734,- 0.5669, 1.568) on the bi-tri branch, obtained via three-mode pseudo-steady approximation.

points in the periodic branches bifurcating at H B 2 , while the pseudo-steady AIFs avoid this artifact. The fate of the periodic solutions for the lower-dimensional traditional Galerkin methods is not clear. While the three-mode traditional approximation fails to find all but a vestige of the symmetric branch bifurcating from the positive bimodal branch, both the four-mode and fivemode truncations yield the primary Hopf bifurcation at HB1, asymmetric limit cycles, and the beginnings of the period doubling sequence of the oscillatory branch bifurcating at HB 1 off the bimodal branch. In addition the four-mode traditional approximation apparently has at least one branch ending in a Sil'nikov loop, and perhaps more. The computational difficulty in tracing the periodic branches, especially around the turning points in a Sil'nikov corkscrew forces us to leave open the existence of those missing. The fivemode traditional approximation exhibits a more definite small distortion: the periodic branch bifurcates from the positive bimodal branch at HB 1, but now does so subcritically, quickly turning back in the correct direction at a spurious turning point. This artifact is interesting in that it was not present for the four-mode traditional approximation, providing a feature of the dynamic behavior

We now turn to a side effect of the AIFs employed here. As mentioned in section 5.1, the nonlinear Galerkin methods failed to find the giant branch within the interval 36 < a < 64. Considering that the bi-tri branches for the pseudosteady AIF bifurcated from the trimodal branches at an inflated value of a, it is reasonable to expect that the bifurcation of the giant branches may also be shifted. In figs. 5a, 5b and 5c we show 300

~

.

Galerkin3modes(a)

150

300 OC

o Z

15C

Z

_o I--_.J

o

CO

0 300

(c)

Pseudo-Steady 3modes 150

iO0

Of ".--,,.-

200

300

Fig. 5. Continuation of the negative bimodal branch for large a for (a) Euler-Galerkin, (h) pseudo-steady If, (c) pseudosteady methods.

56

M.S. Jolly et al. / Inertial manifolds for the Kuramoto-Sivashinsky equation

the continuation of the negative bimodal branch for the Euler-Galerkin, pseudo-steady II, and pseudo-steady AIFs for much higher values of a. While all three methods produce branches on which the solution norm is unbounded, as in the giant branch of the full KSE, they all cause the branches to turn back, and apparently end with a vertical asymptote at a = 0. Since in each case any absorbing ball for the corresponding dynamical system must include the giant solution, the radius of the ball will grow as a decreases, in contrast to the case of the full KSE, which is dissipative. This is also unlike the traditional Galerkin approximations in which the ball suddenly exploded with a steady state bifurcation of the trivial solution. Yet one can show, following Nicolaenko et al. [48], that there exists a*(m) which is an increasing function of m, such that the traditional Galerkin approximation of order m is dissipative for all values of a which are smaller than or* (see ref. [33]). The nonlinear Galerkin case may be explained by the fact that the function representing the AIM is a polynomial of finite degree which, by its growth at infinity, allows an artificial balance of the modes to create a giant steady state at small or. One remedy for this would be to truncate in a smooth fashion the "nonlinear" term, -A1/2u + B(u, u), outside a ball of twice the radius of the absorbing ball, as is done in existence proofs of inertial manifolds (see e.g. refs. [10, 11, 19, 39]). This modification may introduce new dynamic behavior locally, but would not affect the dynamics within the absorbing ball. We investigate this remedy and others in a separate work [33].

6. Conclusions

The computational results here indicate a strong correspondence between the estimates in table 1 and the qualitative performance of the AIFs tested, even though we use such a low number of modes (m = 3) and the rigorous error estimates are asymptotic as m ~ ~. In general,

the dynamics of the nonlinear Galerkin methods seem comparable to that of the traditional Galerkin method using up to twice as many modes. In particular the pseudo-steady AIF, which is based on an AIM passing through the steady states, accurately captures time-periodic behavior, not only locally near Hopf bifurcations, but globally, to the conclusion in Sil'nikov loops. The improvement of the error estimate of the pseudo-steady AIF over the Euler-Galerkin was manifested in two ways: the capture of the tri-bi branch of steady states, and the avoidance of the artificial turning point in fig. 2b. On the other hand, the capture of the tri-bi branch also distinguished the pseudo-steady from the pseudosteady II AIF, which is obtained from the former approximation by neglecting a term estimated to be smaller than the total error. This indicates a sensitivity to certain terms that our measurement of error may not detect. While this work demonstrates the effectiveness of these nonlinear Galerkin methods for the KSE, the potential for computational savings in a large-scale application has yet to be realized. Most of the computational effort in any discretization of the KSE is concentrated on evaluating the nonlinear term. Since the nonlinear Galerkin methods involve the composition of the nonlinearity, direct implementation of the equations for these AIFs will be more costly per mode. For example, in the case of a single-iterate method such as the Euler-Galerkin method, computing the equations for three modes would require the same order of work as the first three equations of the six-mode traditional Galerkin truncation. Nevertheless, the fact that less equations are eventually used can be a distinct advantage in long semi-implicit or implicit integrations, as well as in stability and bifurcation analysis, in which calculations using a smaller Jacobian matrix become a significant factor. Moreover, in the particular case of the KSE over the parameter range of interest here, being able to investigate certain global bifurcations in a three dimensional phase space may prove to be a crucial advantage.

57

M.S. Jolly et aL / Inertial manifolds for the Kuramoto-Sivashinsky equation

Acknowledgements We thank one of the referees for a number of helpful suggestions, which included the addition of the appendix. This work was supported in part by NSF grants CBT-8707090, EET-8717787, DMS 881-0684, DOE Grant DE-FG02-86ER25020, the US Army Research Office through the Mathematical Sciences Institute of Cornell University, and the Packard Foundation.

We denote the complex coefficients in (A.2) by Yk, and using (A.1), rewrite them in terms of the sine coefficients as -1 ~/k

=

--

1i j

+ E

k-1 - j b - j j b k - j + E jbjbk-j m j=l

-

bjbjk,

j=k+l

Appendix

where bj = 0 for j < 1 and for j > m. The m-mode traditional Galerkin approximation is then given by

An m-mode traditional Galerkin approximation amounts to substituting the truncated Fourier expansion

db~ dt = ( - 4 k 4 + akE)bk - afl'~, 1 < k < m,

m

u(x,t)=

(A.3)

Y'~ Uk(t) e ikx where

k=-m

into (KS~) and gathering terms. The only effort comes when evaluating the bilinear term B(u, u ) = u au/ax. The restriction to the space of odd functions requires that the complex coefficients u k be purely imaginary. Setting

fl'~( b 1. . . . . bin) = 2iy k m

l ~_,jbj[bk+j+sgn(k-j)blk_il]

(A.4)

j=l

In particular for m = 3 we have

Uk = -- ½ib k for l_
for - m _
-- 0

otherwise,

dbl - -dt = ( - 4 + a ) b 1+ l a ( b 2 b 3 + b i b 2 ) , (A.1)

one has the sine representation

db3 dt = ( - 3 2 4 + 9 a ) b 3 - ~Otblb 2.

m

u ( x , t ) = Y'. bk(t ) sin(kx). k=l Yet it is easier to multiply complex polynomials to obtain

u

dt = ( - 4 k 4 + a k 2 ) b k

~_, k= --2m

We can express single iterate nonlinear Galerkin methods with the composition of the expression in (A.4) which yields db k

au(x, t) ax =

db2 1 2 - dt - = ( - 6 4 + 4 a ) b 2 + a ( - ~ b I +bib3) ,

j u j ( t ) Uk_y(t m

e ikx.

(A.2)

-- ° t f l 2 m ( b l . . . . .

1
bin,

~m+ 1. . . . . ¢~2m), (A.5)

M.S. Jolly et al. / Inertial manifolds for the Kuramoto-Sivashinsky equation

58

where for m + 1 < j < 2m, we define

is the same as that in (A.3), up to quadratic terms, with additional terms up to degree eight.

ot ta?mgb bm,O . . . . 0), 4'j=4')=- - 4 j 4 ~ J ~, 1. . . . ,

for q~a = 4)1, References

4 ' Y = ~ ' - - 1 f f ~ 4 j 4ti2m(bl . . . . ' b m ' O ' " " O )

for ~a = ~Tfl,r,

(A.6)

with ~1 and ~l,z as in section 4, only adapted for (KS~). The effect in (A.5) is to leave the quadratic terms the same as in (A.3), while adding cubic and quartic terms, which are also quadratic and cubic in the parameter a, respectively. This is most easily seen by the reconstruction of the bilinear function. Let b = (b 1. . . . . bm), 4' = (4'm+ 1. . . . ,4'2m ) and m

B~'( b;b') = 1 y , jb][bk+y + s g n ( k _ j ) b l k _ j l ] . j=l It is easy to see that B m ( b ; b ) = expanding to obtain

tim(b).

After

ti2m( b, 4' ) = B2m( b,O; b,O) -.bB2m(o,4''~b,O)

+ BZm(b,0;0,4,) -1--e2m(0, 4,; 0, 4,), we note that B2m(b,O; b,0) = tim(b, b). The new cubic terms then are introduced by B2m(0, 4,, b, 0) and B2m(b,O;0,4,), while the quartic terms are given by B2m(0, 4';0, 4'). The situation is similar for the second iterate methods. For example, the m-mode pseudo-steady A.IF is given by dbk dt

= (-4k4

+ak2)bk

- atti4m( bl . . . . . bm, 4,m+ l, . . ., 4,4m), 1 < k < m,

(A.7)

where for m + 1 < j _< 4m, we define 4,y---- ¼ a [ - - j - 2 ~ 1 - j - 4 t i ; m ( b , 4 , 1 , 0 , . . . , O ) ] ,

with 4,1 as in (A.6). Again the vector field in (A.7)

[1] C.J. Amick and J.B. McLeod, A singular perturbation problem in needle crystals, Arch. Rational Mech. Anal. 109 (1990) 139-171. [2] D. Armbruster, J. Guckenheimer and P.J. Holmes, Kuramoto-Sivashinsky dynamics on the center-unstable manifold, SIAM J. Appi. Math. 49 (1989) 676-691. [3] J.E. Billotti and J.P. LaSalle, Dissipative periodic processes, Bull. Am. Math. Soc. 77 (1971) 1082-1088. [4] H.S. Brown, Ph.D. Thesis, Princeton University (1990). [5] H.-C. Chang, Travelling waves on fluid interfaces: normal form analysis of the Kuramoto-Sivashinsky equation, Phys. Fluids 29 (1986) 3142-3147. [6] L.-H. Chen and H.-C. Chang, Nonlinear waves on thin film surfaces. II. Bifurcation analyses of the long-wave equation, Chem. Eng. Sci. 41 (1986) 2477-2486. [7] P. Constantin, in: Proceedings of the AMS/SIAM Summer Conference on the Connection between Infinite and Finite Dimensional Dynamical Systems, Boulder, CO, 1987, eds. C. Foias, B. Nicolaenko and R. T6mam, Contemp. Math. 99 (1987). [8] P. Constantin and C. Foias, Global Lyapunov exponents, Kaplan-Yorke formulas and the dimension of the attractor for 2D Navier-Stokes equations, Comm. Pure Appl. Math. 38 (1985) 1-27. [9] P. Constantin and C. Foias, Navier-Stokes Equations (The University of Chicago Press, Chicago, 1988). [10] P. Constantin, C. Foias, B. Nicolaenko and R. T~mam, Integral Manifolds and Inertial Manifolds for Dissipative Partial Differential Equations, Applied Mathematics Sciences, No. 70 (Springer, Berlin, 1988). [11] P. Constantin, C. Foias, B. Nicolaenko and R. T6mam, Spectral barriers and inertial manifolds for dissipative partial differential equations, J. Dynamics Differential Eqs. 1 (1989) 45-73. [12] P. Constantin, C. Foias and R. T~mam, Attractors Representing Turbulent Flows, Memoir AMS, No. 314 (1985). [13] E.J. Doedel, AUTO: a program for the bifurcation analysis of autonomous systems, Cong. Num. 30 (1981) 265-285. [14] C.R. Doering, J.D. Gibbon, D.D. Holm and B. Nicolaenko, Low dimensional behavior in the complex Ginzburg-Landau equation, Nonlinearity 1 (1988) 279-309. [15] E. Fabes, M. Luskin and G.R. Sell, Construction of inertial manifolds by elliptic regularization, IMA Preprint No. 459 (1988). [16] C. Foias, M.S. Jolly, I.G. Kevrekidis, G.R. Sell and E.S. Titi, On the computation of inertial manifolds, Phys. Lett. A 131 (1988) 433-436.

M.S. Jolly et al. /Inertial manifolds for the Kuramoto-Sivashinsky equation [17] C. Foias, O. Manley and R. T6mam, Modelling of the interaction of small and large eddies in two dimensional turbulent flows, Math. Modelling Numerical Anal. 22 (1988) 93-118. [18] C. Foias, B. Nicolaenko, G.R. Sell and R. T6mam, Inertial manifolds for the Kuramoto-Sivashinsky equation and an estimate of their lowest dimensions, J. Math. Pures Appl. 67 (1988) 197-226. [19] C. Foias, G.R. Sell and R. T6mam, Inertial manifolds for nonlinear evolutionary equations, J. Differential Eqs. 73 (1988) 309-353. [20] C. Foias, G.R. Sell and E.S. Titi, Exponential tracking and approximation of inertial manifolds for dissipative equations, J. Dynamics Differential Eqs. 1 (1989) 199-224. [21] C. Foias and R. T6mam, Remarques sur les 6quations de Navier-Stokes stationnaires et les ph6nom~nes successifs de bifurcation, Annali Scuola Norm. Sup.-Pisa, Ser. IV, 5 (1978) 29-63. [22] C. Foias and R. T6mam, Some analytic and geometric properties of the solutions of the Navier-Stokes equations, J. Math. Pures Appl. 58 (1979) 339-368. [23] C. Foias and R. T6mam, The algebraic approximation of attractors: The finite dimensional case, Physica D 32 (1988) 163-182. [24] J.M. Ghidaglia and B. H&on, Dimension of the attractor associated to the Ginzburg-Landau equation, Physica D 28 (1987) 282-304. [25] P. Glendinning and C. Sparrow, Local and global behavior near homoclinic orbits, J. Stat. Phys. 35 (1984) 645-696. [26] M. Graham, P. Steen and E.S. Titi, Dynamics of B6nard convection in porous media: computations with approximate inertial manifolds, in preparation. [27] J.M. Green and J.-S. Kim, The steady states of the Kuramoto-Sivashinsky equation, Physica D 33 (1988) 99-120. [28] J.K. Hale, Asymptotic behavior of dissipative systems, Math. Surveys and Monographs 25 (AMS, Providence, RI, 1988). [29] D.B. Henry, Geometric Theory of Semilinear Parabolic Equations, Lecture Notes in Mathematics, No. 840 (Springer, New York, 1981). [30] J.M. Hyman and B. Nicolaenko, The KuramotoSivashinsky equation: A bridge between PDEs and dynamical systems, Physica D 18 (1986) 113-126. [31] J.M. Hyman and B. Nicolaenko and S. Zaleski, Order and Complexity in the Kuramoto-Sivashinsky model of weakly turbulent interfaces, Physica D 23 (1986) 265-292. [32] M.S. Jolly, Explicit construction of an inertial manifold for a reaction diffusion equation, J. Differential Eqs. 78 (1989) 220-261. [33] M.S. Jolly, I.G. Kevrekidis and E.S. Titi, Preserving dissipation in approximate inertial forms, Los Alamos Preprint LA-UR 89-3230 (1989), J. Dynamics Differential Eqs., in press. [34] I.G. Kevrekidis and M.S. Jolly, On the use of interactive graphics in the numerical study of chemical dynamics,

59

paper No. 22c, presented to the 1987 Annual AIChE Meeting, New York. [35] I.G. Kevrekidis, B. Nicolaenko and C. Scovei, Back in the saddle again: a computer assisted study of the Kuramoto-Sivashinsky equation, SIAM J. Appl. Math., in press. [36] R.E. LaQuey, S.M. Mahajan, P.H. Rutherford and W.M. Tang, Nonlinear saturation of the trapped-ion mode, Phys. Rev. Lett. 34 (1975) 391-394. [37] J.-L. Lions, Quelques M&hodes de R~solution des Probl~mes aux Limites Non Lin6aires (Dunod, Paris, 1969). [38] J. Mallet-Paret, Negatively invariant sets of compact maps and an extension of a theorem of Cartwright, J. Differential Eqs. 22 (1976) 331-348. [39] J. Mallet-Paret and G.R. Sell, Inertial manifolds for reaction diffusion equations in higher space dimensions, J. Amer. Math. Soc. 1 (1988) 805-866. [40] R. Mafi~, On the dimension of the compact invariant sets of certain nonlinear maps, Lecture Notes in Mathematics, No. 898 (Springer, New York, 1981) pp. 230-242. [41] M. Marion, Approximate inertial manifolds for the pattern formation Cahn-Hilliard equations, Math. Modelling Numer. Anal. 23 (1989) 463-488. [42] M. Marion, Approximate inertial manifolds for reaction diffusion equations in high space dimension, J. Dynamics Differential Eqs. 1 (1989) 245-267. [43] M. Marion and R. T6mam, Nonlinear Galerkin methods, SIAM J. Numer. Anal. 26 (1989) 1139-1157. [44] G. M6tivier, l~tude asymptotique des valeurs propres et de la fonction spectral de probl~mes aux limites (These, Universit6 de Nice, 1976). [45] D. Michelson, Steady solutions of the KuramotoSivashinsky equations, Physica D 19 (1986) 89-111. [46] B. Nicolaenko, in: Proceedings of the AMS/SIAM Summer Conference on the Connection between Infinite and Finite Dimensional Dynamical Systems, Boulder, CO, 1987, eds. C. Foias, B. Nicolaenko and R. T6mam, Contemp. Math. 99 (1987). [47] B. Nicolaenko and B. Scheurer, Remarks on the Kuramoto-Sivashinsky equation, Physica D 12 (1984) 391-395. [48] B. Nicolaenko, B. Scheurer and R. T6mam, Some global dynamical properties of the Kuramoto-Sivashinsky equation: Nonlinear stability and attractors, Physica D 16 (1985) 155-183. [49] B. Nicolaenko, B. Scheurer and R. T6mam, Some global dynamical properties of a class of pattern formation equations, Commun. Partial Differential Eqs. 14 (1989) 245-297. [50] C. Scovel, I.G. Kevrekidis and B. Nicolaenko, Scaling laws and the prediction of bifurcations in systems modeling pattern formation, Phys. Lett. A 130 (1988) 73-80. [51] G.R. Sell, Approximation dynamics: hyperbolic sets and inertial manifolds, University of Minnesota Supercomputer Institute, Preprint NO. 89/39 (1989). [52] M. Taboada, Finite dimensional asymptotic behavior for the Swifl-Hohenberg model of convection, Nonlinear

60

M.S. Jolly et al. / Inertial manifolds for the Kuramoto-Sivashinsky equation

Analysis TMA 14 (1990) 43-54. [53] E. Tadmor, The weU-posedness of the KuramotoSivashinsky equation, SIAM J. Math. Anal. 17 (1986) 884-893. [54] R. T6mam, Navier-Stokes Equations and Nonlinear Functional Analysis, CBMS Regional Conference Series, No. 41, SIAM, Philadelphia (1983). [55] R. T6mam, Vari6t6s inertielles approximatives pour les 6quations de Navier-Stokes bidimensionnelles, C. R. Acad. Sci. Paris, Serie II, 306 (1988) 399-402. [56] R. T6mam, Infinite Dimensional Dynamical Systems in Mechanics and Physics (Springer, New York, 1988). [57] R. T6mam, Induced trajectories and approximate inertial manifolds, Math. Modelling Num. Anal. 23 (1989)

541-561. [58] E.S. Titi, Une vari6t6 approximante de l'attracteur universel des 6quations de Navier-Stokes, non lin6aire, de dimension finie, Compt. Rend. Acad. Sci. Paris I 307 (1988) 383-385. [59] E.S. Titi, On approximate inertial manifolds to the Navier-Stokes equations, MSI Preprint 88-119; J. Math. Anal. Appl. 149 (1990), to appear. [60] W.C. Troy, The existence of steady solutions of the Kuramoto-Sivashinsky equation, J. Differential Eqs. 82 (1989) 269-313. [61] S. Wiggins, Global Bifurcations and Chaos, Applied Mathematics Sciences, No. 73 (Springer, New York, 1988).