This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
0, and we define ^ = I«Mitl,2toH/2, o •'n •'n { g i ( v - >) dT, = 0 at the trailing edge T.E. of B. = -
(4-34)
where Xi+i/i denotes the characteristic function of the open segment ]M f , Mi+ j[. Then 4>h > 0 a.e. on F,
lim || ^ - 01|L . ( r ) = 0.
(4.35)
fc->0
Let us consider a sequence (vh)h,vheK^,'ih,
such that
lim vh = v weakly in V. (4.36) 2 From (4.36) (see Necas [1]) it follows that limft_»0 yvh = yv strongly in L (F). This in turn implies that lim f y i v k d r =
\yv
(4.37)
From Simpson's rule it follows that \yvh<j>hdT = - £ lAf.Mj+'.WAf
J f o M ) + 4vh(Mi+ll2) Vi)»eKj ! ,
+ vh(Mi+J]
V^eC°(F),
From (4.37) and (4.38) we obtain (* yv
V (j) e C°(F),
•Jr
which implies yy > 0 a.e. on F. This proves (i).
4> > 0,
> 0,
(/> > 0.
(4.38)
4. A Third Example of EVI of the First Kind: A Simplified Signorini Problem
63
Verification of (ii). From Lemma 4.2, it is natural to take x = K n C°°(Q). Define r\: H\a) n C°(fi) -^ Fj by r\veV\,
y v e H\a)
t*v{P) = i>(P),
n
CiH),
V P e S j , k = 1, 2.
(4.39)
On one hand, under the assumptions made on 9~h, we have (see Strang and Fix [1]) llrju - »||F < Chk\\v\\Hk,Hii),
V i; £ C M (a),
k = 1, 2,
(4.40)
with C independent of /i and y. This implies Z)
fc = 1,2.
(4.41)
On the other hand, it is obvious that r\v e Kj, V i; e K n C0(Q), so that rju e Kj, V « e z, k = 1, 2. In conclusion, with the above % and rkh, (ii) is satisfied. •
Remark 4.8. For error estimates in the approximation of (4.5) by piecewise linear finite elements, it has been shown by Brezzi, Hager, and Raviart [1] that we have
assuming reasonable smoothness for u on Q.
4.6. Iterative methods for solving the discrete problem We shall briefly describe two types of methods which seem to be appropriate for solving the approximate problems of Sec. 4.4. 4.6.1. Solution by an over-relaxation method The approximate problems (P%) are, for k = 1,2, equivalent to the quadratic programming problems described in Remark 4.6. By virtue of the properties of K\ (see Sec. 4.4.1), for the solution of (P%), we can use the over-relaxation method with projection, which has already been used in Sec. 2.8 to solve the approximate obstacle problem and which is described in Chapter V, Sec. 5. From the properties of our problem, the method will converge, provided 0 < co < 2. 4.6.2. Solution by a duality method We first consider the continuous case. Let us define a Lagrangian J? by
-L(v)-
[qyvdT, •IT
(4.42)
64
II Application of the Finite Element Method
and let A be the positive cone of L 2 (F), i.e., A = {q € L2(T), q>0 a.e. on F}. Then we have:
Theorem 4.4. Let L(v) = | n fv dx + j r gyv dT with f and g sufficiently smooth. Suppose that the solution u of (4.5) and (4.6) belongs to H 2 (O); then {u, du/dn — g} is the unique saddle point of£? over Hl(Q) x A. PROOF. We divide the proof into two parts. In the first part we will show that {u, du/dn — g] is a saddle point of <£ over Hi(Q) x A, and in the second part we will prove uniqueness. (1) Let p = du/dn — g. From the definition of a saddle point, we have to prove that &(u, q) < y{u, p) < y(v, p), 2
V{v,q}eV
l2
x A, {«, p} e V x A .
(4.43)
2
Since u e H (Q.), we have du/dn e H ' (T) c= L (T) (see Lions and Magenes [1]). Then if g is smooth enough, we have p = du/dn — g e L2(T). From Proposition 4.1 we have du p=
#>0onr, 8n
(4.44) pyu = 0 a.e. on F. This implies that we have {u,p} 6 H1(Q) x A and j r pyu dT = 0. Since yu > 0 on F, we have {qyudT>0,
V q e A.
(4.45)
From (4.44) and (4.45) it follows that &{u, q) = ia(u, u) - L{u) - \ qyu dT < \a(u, u) - L(u) = |a(u, u) - L{u) - [ pyu dT = JS?(«, p),
Vq 6 A
•Jr
which proves the first inequality of (4.43). To prove the second inequality of (4.43), we observe that the solution u* of the minimization problem Se{u", p) < £C(v, p),
V c e H'(fi),
u* e H^O)
(4.46)
is unique and is actually the solution of the linear variational equation a(u*, v) = L(v) + I pyv dT,
V v e H^Q),
u* e H^Q).
(4.47)
Jr Since L(v) = \afv dx + J r gyv dT, u* is actually the solution of the Neumann problem - A M * + u* = / i n f i , du* du — =p + , = - o n F .
4. A Third Example of EVI of the First Kind: A Simplified Signorini Problem
65
Since from Proposition 4.1 we obviously have —AM + u — fin Q, du
du*
y- = -x- on T, on on it follows from the uniqueness property of the Neumann problem (4.48) that u — u*. Using (4.46) and u = u*, we obtain the second inequality in (4.43). This proves that {u, p} is a saddle point of i f over /^(Q) x A. (2) Uniqueness. Let {«*, p*} be a saddle point of if over H^Q) x A. We will show that u* = u,p* = p. From (4.42) and (4.43) it follows that Up-
q)yu dT < 0,
V q e A.
(4.49)
V <j 6 A.
(4.50)
Similarly, we have f (p* - <j)yu* dF < 0,
Taking q = p* (respectively, q = p) in (4.49) (respectively, (4.50)), we obtain (p* - p)y(u* - u) dT < 0. r From the second inequality of (4.43) it follows that u is the solution of a(u, v) = L(v) + (* pyv dT,
(4.51)
V c e H\Q),
u e H'iSi),
(4.52)
V v e H^Q),
u* e H^Q).
(4.53)
and similarly, a(u*,v) = L(v)+
\p*yvdT, •JT
Taking v = u* — u (respectively, v = u — «*) in (4.52) (respectively, (4.53)), we obtain a(u* -u,u*
- M) = [ (p* - p)7(u* - M) dr. •'r
(4.54)
Using the F-ellipticity of a(-, •), from (4.51)-(4.54) it then follows that u* = u and
I (p* - p)y^rfr = o,
VceHHS),
•'r
which implies that p* = p. Hence {u, p] is the unique saddle point of i f over H1(Q) x A.
•
66
II Application of the Finite Element Method
From Theorem 4.4 it follows that we can apply Uzawa's algorithm to solve (4.5) (see Cea [2, Chapter 5] and G.L.T. [1, Chapter 2], [2, Chapter 4, Sec. 3.6], [3]). In the present case this algorithm is written as follows: p° e A is arbitrarily chosen (for instance, p° = 0);
(4.55)
then p" being known, we compute, for n > 0, {un, pn+1} by &{un, pn) < <£{y>, p"),
V c e H^Q),
u" e H\Q),
(4.56) (4.57)
2
2
where PA is the projection operator from L (F) to A in the L (F) norm, and p > 0. From (4.56) it follows that u" is in fact the solution of the Neumann problem -AM" + u" = / i n Q , dif
l)n~
(4.58) r
The projection PA is given by PM
= q+ ,
V<jeL 2 (F).
(4.59)
Since y: H1(Q) -» L 2 (F) is a continuous linear mapping, we have \\yv\\LHr) < \M\ • \\v\\HHn),
V v e H\O).
(4.60)
From Cea and G.L.T., loc. cit., it then follows that lim M" = u strongly in H\Q),
(4.61)
where u is the solution of the problem (4.5) provided that 0 < p < 2/||y||2. Let us give a direct proof for this convergence result. This proof will use the characterization (4.5) given in Proposition 4.1 (even if a(-, •) is not symmetrical, the same result follows). It will be convenient to take (4.56), (4.58) in the following equivalent form: a(u", v) = L(v) + f p"yv dY, Jr
V U G H^Q),
U" e H\C1).
(4.62)
Let u be the solution of (4.5) and p = du/dn — g. From Proposition 4.1 it follows that a(u, v) = L(v) + f pyv dY, Jr
I
(q - p)yu dY > 0,
r
V c £ Hl{Q),
V«eA,
u e H\Si),
p e A.
(4.63)
(4.64)
4. A Third Example of EVI of the First Kind: A Simplified Signorini Problem
67
Relation (4.64) can also be written as (i — P)(P — Pyu — p) dT < 0,
V q e A, p > 0 and p e A,
which is classically equivalent to p=PA(p-pyu).
(4.65)
Let us consider u" = u" - u,
p" = p" - p.
Since J°A is a contraction, from (4.57) and (4.65) we have \\pn+1\\mr)<\\pn-pyan\\^r)-
(4.66)
From (4.66) it follows that n n
p dT - p2||yu"||£2(r).
(4.67)
Taking v = u" in (4.62) and (4.63), we obtain a{u",u")=
\pnyundY. Jr
(4.68)
From (4.67) and (4.68) it then follows that IIPi£i,n - IIP"+1ll£>(r) > P(2 - PIMI2)II«1^(Q)-
(4.69)
2
If 0 < p < 2/||y|| , we observe that the sequence {||p"||i2(n}n is decreasing and hence converges. Therefore we have
lim(||p"||£ 2 ( r ) - ||p"+1 | | £ W = ° n-* GO
so that lim
||M"|| H . (n) = 0.
«~> CO
Since u" = u" — u, we have proved the convergence. Similarly we can solve the approximate problems (P\h), k = 1,2, using the discrete version of algorithm (4.55)-(4.57). We shall limit ourselves to k = 1, since the extension to k = 2 is trivial. Here we use the notations of Sec. 4.1. Assume that yh = EA — t,h has been ordered. Let yh =
68
II Application of the Finite Element Method
We approximate A and 5£ by Aj|= {qh\qh= {;};, 4i > 0} and ^l(vh, qh) = Hvh, vh) - L(vh) + qi+1vh(Mi+1)l
(4.70)
We can prove that JSfjJ has a unique saddle point {uh, ph}, where ph is a F. John-Kuhn-Tucker vector for (T[h) over V\ x A/J and uh is precisely the solution of (P\h). The discrete analogue of (4.55)-(4.57) is then pJeAj, n
^l(u h,
n
P h) < ^l(vh,
p"h),
(4.71) VvheVl,
Pl+Y = ip"-pul(MdT,
u"heVl
Vi, p > 0 .
(4.72) (4.73)
One can prove that if 0 < p < /?, with /? small enough, then limn^ + 00 «^ = wh, where uh is the solution of (P\h). In G.L.T. [3, Chapter 4] one may find numerical applications of the above iterative methods for piecewise linear and piecewise quadratic approximations of (4.5). EXERCISE
4.3. Extend the above considerations to (Pfh).
5. An Example of EVI of the Second Kind: A Simplified Friction Problem 5.1. The continuous problem. Existence and uniqueness results Let Q be a bounded domain of R2 with a smooth boundary F = <9Q. Using the same notations as in Sec. 4, we define V = H\Q), a(u, v) =
VM
Jn
• Vv dx + uv dx, Jn
L(c) = ,»>, j(v) = g We have then the following:
\yv\dT,
feV*, where g > 0.
(5.1) (5.2)
(5.3) (5.4)
5. An Example of EVI of the Second Kind: A Simplified Friction Problem
69
Theorem 5.1. The variational inequality a(u, v-u)+
j(v) - j(u) >L(v-u),
VveV,
ue V,
(5.5)
has a unique solution. PROOF. In order to apply Theorem 4.1 of Chapter I, it is enough to verify thatj(-) is convex, proper, and l.s.c. Actually j(-) is a seminorm on V. Therefore, using the Schwarz inequality in L2(T) and the fact the y e J5?(ff'(fi), L2(F)), we have \j(u) - j(v)\ < \j(u -v)\<
3(meas(r))1/2Hy(t> - «)|| L2(r) < C\\u - v\\v
(5.6)
for some constant C. Hence j( •) is Lipschitz continuous on V, so that j( •) is l.s.c. ;j{ •) is obviously convex and proper. Hence the theorem is proved. •
Remark 5.1. If g = 0, it is easy to prove that (5.5) reduces to the variational equation a(u, v) = L(v),
VceF,
u e V.
This is related to the variational formulation of the Neumann problem. Remark 5.2. Since a( •, •) is symmetrical, the solution u of (5.5) is characterized, using Lemma 4.1 of Chapter 1, as the unique solution of the minimization problem J(u) < J(v),
V v e V, ueV,
(5.7)
where J(v) = \a(v, v) + j(v) - L(v). Remark 5.3. Problem (5.5) (and (5.7)) is the simplified version of a friction problem occurring in elasticity. For this type of problem we refer to Duvaut and Lions [1] and the bibliography therein. EXERCISE
5.1. Let us denote the solution of (5.5) by ug. Then prove that lim ug = u strongly in H J (Q), ff~* + 00
where u is the unique solution of a(u, v) = Hy),
V D G ffJ(Q), u e Hj(Q).
5.2. Regularity of the solution Theorem 5.2 (H. Brezis [3]). If (I is a bounded domain with a smooth boundary and if L(v) = \n fv dx with f e L2(Q), then the solution u of (5.5) is in H2(Q).
70
II Application of the Finite Element Method
5.3. Existence of a multiplier Let us define A by
A = {^ e L\T), \ix{x)\ < 1 a.e. on T). Then we have: Theorem 5.3. The solution u of (5.5) is characterized by the existence of X such that a(u, v) + g f Xyv dT = L(v), Jr X e A,
V v e V,
ueV,
Ayw = | yu | a.e. on F.
(5.8) (5.9)
PROOF. We will prove first that (5.5) implies (5.8) and (5.9). Taking v = 0 and v = 2M in (5.5), we have a(u, u) + j(u) = L(u).
(5.10)
It follows then from (5.5), (5.10) that L(u) - a(u, v) < j(v),
VCEF,
which implies \L(v)
- a(u, v)\ < j(v) = g \ \yv\ dT,
V c e V.
(5.11)
We have H 1 ^ ) = #J(Q) © [^o(fi)]1, where [Hj(Q)]1 is the orthogonal complement of Hl(Q) in H^Q). Since y: [//J(Q)] 1 -» H 1 / 2 (r) is an isomorphism, it follows from (5.11) that L(v) - a(u, v) = l(yv),
MveV,
(5.12)
1I2
where /(•) is a continuous linear functional on H (T). It then follows from (5.11), (5.12) that IWI<0MlL.(n,
V^6H"!(I>
(5.13)
Since H 1 / 2 (r) ez Ll(T), it follows from (5.13) that we can apply to /(•) the HahnBanach Theorem (see, for instance, Yosida [1]) to obtain the existence of -16 L co (r), I A(x) | < 1 a.e. on T such that (5.14) Therefore it follows from (5.12) and (5.14) that a(u, v) + g f Xyv dT = L(v), which proves (5.8).
V v e V,
5. An Example of EVI of the Second Kind: A Simplified Friction Problem
71
Taking v = u in (5.8), we obtain a(u, u) + g \ Xyu dT = L{u). •Jr
Using (5.10) and the above equations, we obtain U\yu\ - Xyu) dT = 0. •>r
(5.15)
Since | X \ < 1 a.e., we have \yu\-
Xyu> 0a.e..
(5.16)
From (5.15) and (5.16) it follows that | yu | = Xyu a.e.. This completes the proof of (5.8) and (5.9). Assuming (5.8) and (5.9), we will show that (5.5) holds. Let {u, X} be a solution of (5.8), (5.9). From (5.8) it follows that a(u, v — u) + g
Xy(v — u) dT = L(v — u),
V v e V,
which can also be written as a(u, v - u ) + g \ XyvdY Jr From (5.9) and (5.17) we obtain a(u, v-u)
- g [Xyu dT = L(v - u ) , •'r
VceK
+ g [xyvdT - g f \yu\ dT = L(v - u), Jr Jr
VoeF.
(5.17)
(5.18)
But since Xyv < | yv \ a.e. in T, it follows from (5.18) that a(u, v - u) + j(v) - j(u) > L(v - u),
VteF.
This proves the characterization.
D
Remark 5.4. Assuming that
L(v)= f fovdx+ Ja
[j\yvdT, Jr
with/o, / i sufficiently smooth, we can express (5.8) by — AM + H = / O in Q
^ + gX = j \ a.e. on r . on From (5.19) it follows that X is unique. EXERCISE 5.2. Prove that X is unique.
(5.19)
72
II Application of the Finite Element Method
5.4. Finite element approximation of (5.5) Let Q be a bounded polygonal domain of R2. The notation used here is largely the same as in Sec. 4.4 of this chapter. 5.4.1. Approximation of V We use the piecewise linear and piecewise quadratic approximations of V = H^Q) described in Section 4.4.1 of this chapter. 5.4.2. Approximation ofj( •) We use the notation of Figure 4.2. Then we approximate _/(•) by
H(vh) = § £ IMji i + ! |(|yvh{Md| + |yvh(Mi+ 01), = lY. \MiMi+l\(\yvh(Mi)\ + |y»»(Afi+1)|),
V vh e Vlh,
(5.20)
+ 4\yvh(Mi+1/2)\
V» t eK ( 2 .
In (5.20) and (5.21) we have Mt e yh and Mi+1/2
(5.21)
e yj,.
Remark 5.5. Clearly (5.20) and (5.21) are obtained from j(-) by using trapezoidal and Simpson's numerical integration formulae, respectively. 5.4.3. The approximate problem For k = 1, 2, problem (5.5) is approximated by fl(«J, t;* - «J)+;£(»,,)-jJ(«!D>M» f c -"ID.
Vi) t eV{,
«j€Vt
(P*2fc)
Then: Proposition 5.1. Problem (P^h) has a unique solution. Remark 5.6. Since a(-, •) is symmetrical, (Pk2h) is equivalent to the nonlinear programming problem Min[_Hv h , vh) + ; > * ) - L(vh)l
(5.22)
Remark 5.7. Using (5.20), (5.21), and (7.1)-(7.4) of Sec. 7 of this chapter, we may express (P^) and (5.22) in a form more suitable for computations.
5. An Example of EVI of the Second Kind: A Simplified Friction Problem
73
5.5. Convergence results Theorem 5.4. Suppose that the angles of 2Th are uniformly bounded below by 60 > 0 as h -> 0; then lim u\ — u strongly in H1(Q),
(5.23)
where u and u\ are, respectively, the solutions of (5.5) and (P2/,) far k = 1, 2. PROOF. TO prove (5.23) it suffices to verify the following properties (see Theorem 6.3 of Chapter I). (i) There exists U c V, U = V and r\: U ->• V\, such that lim r\v = v strongly in V, (ii) Vvk-*v
Vcel/.
weakly in V, then lim Mfh(vh) > j(v).
(iii) \imjkh(rkhv)=j{v),
V veU.
Verification of(i). Since F is Lipschitz continuous, we have (see Necas [1]) C 0 0 ® = H\n).
(5.24)
m
Therefore it is natural to take U = C (U). Define r\ by (4.39) of Theorem 4.3, Chapter II; under the above assumptions on ,fh, it follows from Strang and Fix [1] that \\A.v-v\\v
VoeK,
(5.25)
where C is a constant independent of h and v. This implies (i). Verification of (ii) (1) Case k = 1. We again use the notation of Fig. 4.2. Since the trace of vh restricted to [Mj, Mi+l~\ is affine, it follows that yvh(M) =
,
(\MMi+1\yvh(Mi)
+
\MMi\yvh{Mi+l)),
VvheVl,
VMe[Mi,M i + 1 ].
(5.26)
Since
the convexity of £ -> \£\ implies
I yvh(M) I < - = 4 |MM |M f M| \yvh(Mi+l)\),
VvheVl
¥Me[M,Mitl],
(5.27)
II Application of the Finite Element Method
74 Integrating (5.27) on MtMi+,, (
we obtain i+1
\yvh\dT<—'
which implies that, V vh € F^, we have
f
v- f
Thus we have proved j(vh)
VvheVl
(5.28) 2
Let uft -> f weakly in K Then limA^0 yuh = 71; strongly in L {Y), which implies \imj{Vh)=j{v).
(5.29)
fi->0
It then follows from (5.28) and (5.29) that lim infft^0 jl(vh) > j(v), which proves (ii) iffc = 1. (2)Casek = 2. Let us define M i + 1 / 6 , M 1 + 5 / 6 by (see Fig. 5.1)
Then we define qh: C°(r) ^ L°°(r) by (5.30)
where Z , (respectively, Xi+ll2) is the characteristic function of M;_ 1 / 6 MjM i + 1 / 6 (respectively, M i + 1 / 6 A? j + 5 / 6 ). We then have the following obvious properties: lim qh(ji) = n strongly in L co (r),
V n e C0(T),
= g\\qkyvk\\LHr),
.II w , <
w < C 2 ||7^|| L2(r) ,
V «„ e Kft2, V !>* e F,2,
1/2
Figure 5.1
(5.31) (5.32) (5.33)
5. An Example of EVI of the Second Kind: A Simplified Friction Problem
75
where in (5.33), Cj and C 2 are positive constants independent of vh, h, and T (values for C t and C2 may be found in G.L.T. [3, Chapter 4]). Now, taking n e C°(F), we define sh(n) by sh(n) e L°°(F),
s,(/i)|]Mi,M( + lI = MM ; + 1 / 2 ).
(5.34)
Then lim sh(n) = n strongly in L°°(r),
V / J E C°(r),
(5.35)
/i->0
and from Simpson's integration formula, we have \ sh{ii)qhyvhdY = Lh(n)yvhdr,
V /i e C°(r),
Vt.eF,2.
(5.36)
Let t)ft -> u weakly in V, vh e V\, V h; then lim yvh = yu strongly in L2(T).
(5.37)
On the one hand, it follows from (5.33) that > < C,
(5.38)
where C is independent of h. On the other hand, (5.31), (5.35)-(5.38) imply that lim f sh(fi)qhyvh dT = \ nyv dT, h^o •'r
V fi e C°(T).
(5.39)
Jr
In turn, (5.38) and (5.39) imply that lim qhyvh = v weakly in L2(T).
(5.40)
Since the functional n -^ !|/u||Li(n) is convex and continuous on 1/(1"), it follows from (5.40) that lim inf \\qhyvh\\LHr) > ||yt!||L.(n.
(5.41)
/i^O
Combining (5.41) with (5.32), we obtain lim inffc^0 j2(vh) > j(v) which proves (ii) when k = 2. Verification of (iii). Let veU = C^fi). From (5.25) and from the uniform continuity of yv on F, it follows almost immediately that \imfh(rkhvh)=j(v),
/c = l,2.
Since conditions (i), (ii), and (iii) are satisfied, the strong convergence of uh to u follows from the Theorem 6.2 of Chapter I. D Remark 5.8. It is proved in G.L.T. [3, Chapter 4] that vh -* v weakly in V, vh e V\, implies l i m ^ 0 j£Oj,) = j(v), fe = 1, 2.
76
II Application of the Finite Element Method
Since the proof of this result is rather technical we have used a simpler approach in this book, from which it follows that lim Mfh(vh) > j(v),
k = 1, 2.
As we have seen before, this result was sufficient for proving Theorem 5.4. 5.6. Iterative methods for solving (P*fc) In this section we briefly describe some iterative methods which may be useful for solving the approximate problems (P|fc). 5.6.1. Solution of(Pk2h) by relaxation methods It follows from (5.20)-(5.22) (see Remark 5.6) that (Pk2h), k= 1,2, are particular cases of Min f(v), (5.42) where, with v = {vu ..., vN), N
f(v) = \{Av, v) - ( b , v ) + X on I v, I •
(5.43)
In (5.43), (•, •) denotes the usual inner product of UN, A is a N x JV symmetrical positive-definite matrix, and at > 0, V i = 1 , . . . , N. It then follows from Chapter V, Sec. 5, Cea [2, Chapter 4], Cea and Glowinski [1], and G.L.T. [3, Chapter 2] that we can use a relaxation method for solving (5.42). Actually, from the computations we have done, it appears that the introduction of an over-relaxation parameter at, w > 1, will increase the speed of convergence. Finally the algorithm we used is the following: u° arbitrarily given in UN;
(5-44)
then for i = 1, 2 , . . . , N, f(u\+ \...,
utl,
M?+ \ «?+1, • • •) < / «
\ • • •, utl,
vh u"i+ u ...),
VVi e U, (5.45),
If co = 1, (5.44)-(5.46) reduces to a relaxation method. Numerical solutions of (5.5) using (5.44)-(5.46) are given in G.L.T. [2, Chapter 4], [3, Chapter 4]. Remark 5.9. If a; > 0, u"+1 is the solution of a single variable nondifferentiable minimization problem which can be exactly solved by simple calculations. EXERCISE
5.3. Express u"+1 as a function of A, b, u", un+1.
5. An Example of EVI of the Second Kind: A Simplified Friction Problem
77
5.6.2. Solution o/XP^) by a duality method We first examine the continuous case. Define a Lagrangian £C:H1(Q) x L2(T) - D3 by
Se(v, p) = fr(v, v) - L(v) + g [nyv dY.
(5.47)
Then, using the notation of Sec. 5.3, the following theorem follows from Theorem 5.3. Theorem 5.5. Let {u, X) be a solution of (5.8), (5.9); then {u, X} is the unique saddle point of ^ over H1(Q) x A. EXERCISE
5.4. Prove Theorem 5.5.
From Theorem 5.5 it follows that to solve (5.5) we can use the following Uzawa algorithm: X° e A arbitrarily chosen (for instance, X° = 0); +1
then by induction, knowing X", we compute u" and X" J?(w", X") < £?{v, X"), n+1
X
V » 6 H\O),
= PA(X" + pgyu"),
(5.48)
by U" e V,
p > 0.
(5.49) (5.50)
The minimization problem (5.49) is actually equivalent to the Neumann variational problem a(u", v) = L(v) - g \ X"yv dY,
V u e H\n\
u" e H\O).
(5.51)
In (5.50), P A is the projection operator from L2(Y) to A in the L 2 -norm; then PA(A<) = Sup(-l,Inf(l,AO),
V^eL2(r).
(5.52)
Using Cea [2] and G.L.T. [1, Chapter 2], [3, Chapter 2], it follows that for 0 < p < 2/g2\\y\\2 we have lim u" = u strongly in H1(Q);
u is solution of (5.5), (5.7).
As in Sec. 4.6.2, a direct proof of the convergence of (5.48)-(5.50) can be given; however, it will use the results of Theorem 5.3. 5.5. Using Theorem 5.3, give a direct proof of the convergence of (5.48)-(5.5O).
EXERCISE
78
II Application of the Finite Element Method
The adaptation of (5.48)-(5.5O) to the discrete problems (P|A), k = 1, 2, is straightforward (see G.L.T. [3, Chapter 4]), since it is a simple variant of the discrete algorithm described in Sec. 4.6.2. EXERCISE
5.6. Study the discrete analogues of (5.48)-(5.5O) related to (P^),
k = 1, 2.
6. A Second Example of EVI of the Second Kind: The Flow of a Viscous Plastic Fluid in a Pipe Most of the material in this section can be found in G.L.T. [2, Chapter 5], [3, Chapter 5] and in Glowinski [1], [3].
6.1. The continuous problem. Existence and uniqueness results
Let Q be a bounded domain of U2 with a smooth boundary T. We define V = Hj(fi), a(u, v) =
VM • Vv dx,
L(v) = a\v}, j(v)=
feV*,
\\Vv\dx. Jn
Let n, g be two positive parameters; then: Theorem 6 . 1 . The variational
inequality
Ha(u, v - u) + gj(v) - gj(u) > L(v has a unique
u),
VueF,
ueV
(6.1)
solution.
PROOF. I n o r d e r t o a p p l y T h e o r e m 4.1 o f C h a p t e r I, w e o n l y h a v e t o verify t h a t }{•) is c o n v e x , p r o p e r , a n d l.s.c.. It is o b v i o u s t h a t j(•) is c o n v e x a n d p r o p e r . L e t u,veV; then \j(v) - j(u)\ < y m e a s ^ H u - v\\v; hence XO is l.s.c. This proves the theorem. EXERCISE 6.1. Prove that j(-) is a n o r m on V.
(6.2)
•
6. A Second Example of EVI of the Second Kind: Flow of Viscous Plastic Fluid in a Pipe
79
Remark 6.1. If we take g = 0 in (6.1), we recover the variational formulation of the Dirichlet problem — fiAu = f in Q, u = 0 on F. Remark 6.2. Since a{-, •) is symmetrical, the solution u of (6.1) is characterized (using Lemma 4.1 of Chapter I) as the unique solution of the minimization problem J(u) < J(v),
VceF,
ueV,
(6.3)
where J(v) = (ji/2)a(v, v) + gj(v) — L(v).
6.2. Physical motivation If L(v) = C\nv dx (for instance, C > 0), as proved in Duvaut and Lions [1, Chapter 6], (6.1) models the laminar stationary flow of a Bingham fluid in a cylindrical pipe of cross section Q, u(x) being the velocity at x e Q (we refer to Prager [1], Germain [1], and Duvaut and Lions [1, Chapter 6] for the definition of Bingham fluids). The constant C is the linear decay of pressure and H, g are, respectively, the viscosity and plasticity yield of the fluid. The above medium behaves like a viscous fluid (of viscosity /J.) in Q + = {xefi, | and like a rigid medium in fi° = {x e Q, VM(X) = 0}. We refer to Mossolov and Miasnikov [1], [2], [3] for a detailed study of the properties of Q + and Q°. We observe that (6.1) also appears as a free boundary problem.
6.3. Regularity properties Theorem 6.2 (H. Brezis [4]). / / L(v) = \Qfv dx,feL2(Q), of (6.1) satisfies
then the solution u
u e V n H 2 (Q), and ifQ. is convex, we have
IMI
<
(6-4)
80
II Application of the Finite Element Method
6.4. Further properties
Let us denote by a the following quantity: a -
inf
J
^
(6.5)
v
Then a > 0. From this we derive the following Proposition 6.1. Let u be the solution of (6.1) and fe L°°(Q); then u = 0 if
tl/IL-w) ^ 0«-
(6-6)
PROOF. By taking v = 0, v = 2« in (6.1), we obtain /«J(M, U) + gj(u) =
fu dx.
(6.7)
•>n It then follows from (6.5) and from \nfu dx < ||/||L=o(n)l|M|!L!(fi)tnat Ha(u, u) + (got - ||/||f.=o(n))||M||t,(n) < 0.
(6.8)
If / obeys (6.6), it then follows from (6.8) that u = 0.
•
We also have:
Proposition 6.2. Let u be the solution of (6.1) and f e L"(Q), p > 1. Then if f > 0, we have u > 0. EXERCISE
6.2. Prove Proposition 6.2.
Proposition 6.3. Let u be the solution of (6.2) and f = c, a constant. Then u = 0 if and only ifc < go. and u / 0 if c > ga. EXERCISE
6.3. Prove Proposition 6.3.
Proposition 6.4. Let u be the solution of (6.1) and feL2(Q);
then u = 0 if
where -
inf
PROOF. From a result of L. Nirenberg and M. Strauss (cf. Strauss [1]), it follows that j8>0. By taking v = 0 and v = u in (6.1), we obtain Ha(u,u) + gj(u)=
f fu dx. Jo
(6.10)
6. A Second Example of EVI of the Second Kind: Flow of Viscous Plastic Fluid in a Pipe an
Using \afu dx < \\f\\L2(n)\\u\\L2m
81
d /? > 0, we obtain
im(u, u) + (gp - ||/|| L 2 (n) )||«|| L2(Q) < 0.
(6.11)
Hence, if/satisfies (6.9), from (6.11) we have a(u, u) = 0. This implies u = 0, hence, the proposition is proved. •
6.5. Exact solutions 6.5.1. Example 1 We take Q = ]0, 1[ a n d / = c, c a positive constant. In this case the solution of H Jo
u(v' - u')dx + g \ \v'\dx - g \ \u'\ dx > c \ (v - u)dx, Jo Jo Jo VveH0(Q),
ueH0(Q), (6.12)
(where v' = dv/dx) is given by u = 0ifg>C-.
(6.13)
If g < c/2,
= —
-I
2\i \2
if
cj
< x <- + -, 2
c
2
(6.14)
c
u(x) = ^ - x ( l - x ) - - ( 1 - x) if- + - < JC < 1. We observe that if g < c/2, u e #J(Q) n PF2> 0O(Q), but u ^ H 3 (Q). 6.5.2. Example 2 Let Q = {x|xf + x\ < R2}, j ' — C, C a positive constant. Then the solution of (6.1) is given by w= 0 ifg(>—,
(6.15)
If g < CR/2, then ifR'
«W=|^r^ llT(* + *')-2ff|
ifO
82
II Application of the Finite Element Method
where
We also observe that tig < CR/2, then u e H^Q) n W2' °°(Q), but u $ H3(Q) (we actually have u e HS(Q), s < §). EXERCISE 6.4. Verify that (6.13), (6.14) and (6.15), (6.16) are solutions of (6.12)
and (6.1), respectively.
6.6. Existence of multipliers Let us define A by A = {q\q e L2(Q) x L2(Q), \q{x)\ < 1 a.e.} with \q{x)\ = yfq\{x) + q22(x); then we have: Theorem 6.3. The solution u of (6.1) is characterized by the existence of p such that Lia(u,v) + g f p • VP dx = ,«>, p-V« = |Vw|a.e.,
V»eK,
u 6 V,
p e A.
(6.17) (6.18)
PROOF. There are classical proofs of the above theorem using Min-Max or HahnBanach Theorems (see, for instance, Cea [2, Chapter 5], G.L.T. [1, Chapter 1], and Ekeland and Temam [1]). In the sequel following G.L.T. [3, Chapter 5] we shall give an "almost constructive" proof, making use of a regularization technique. (1) We first prove that (6.17), (6.18) imply (6.1). It follows from (6.17) that Ha(u, v — u) + g
p • V(v — u) dx = na(u, v — u) + g = ,[)-«>,
p - V v dx — g
Vt>eK
p • V « dx (6.19)
It follows from (6.18) that [p-Vudx=
f \Vu\dx,
(6.20)
and from the definition of A that \ p-Vvdx< •In
( \p\\Wv\dx< Jo
\ \Vv\dx, Jfi
VueF.
(6.21)
6. A Second Example of EVI of the Second Kind: Flow of Viscous Plastic Fluid in a Pipe 83 Then, from (6.17), (6.19)-(6.21), we obtain
/M(U, v — u) + gj(v) — gj(u) > , v — u),
VneF, u e V.
Thus (6.17) and (6.18) implies (6.1). (2) Necessity of (6.17) and (6.18). Taking u = 0 and v = 2u in (6.1), we obtain Ha(u, u) + gj(u) = , «>.
(6.22) 2
2
Let e > 0. Regularize j(-) by jE(-) defined by ;,.(«) = Jn^/e + |V«| rfx.Since j £ (-) is convex and continuous on V, the regularized problem lia(uc,v-ut)
+ gje{v)-gjr(ur)>L{v-ur),
V v e V, u£eV,
(6.23)
has a unique solution. Let us show that limE^0 uc = u strongly in V. From (6.1) and (6.23) it follows that lia(uc, u - ur) + gjc{u) - gje(ue) > L(u - uE), Ha(u, uE-
u) + gj(ut) - gj(u) > L(uc - u).
Adding these inequalities we obtain Ha{uc -u,u,-
u) + g(jr(us) - j(uj) < g(jF{u) - j(u)).
(6.24)
From 0 <- Jt2
+ F2 - \ t \ =
< £, ^
VfeIR,
'
it follows that 0 < j£v) - j(v) < e meas(fi), V' v e V, which using (6.24), implies that lia(iiz — u, ut — u) < ge meas(fi), so that / (q V / 2 II«r - u\\v < v m e a s ^ ) - e .
(6.25)
V i From (6.25) we obtain lim us = u strongly in V.
(6.26)
Since j e ( ) is differentiable on V, problem (6.23) is equivalent (see, e.g., Cea [1]) to the following nonlinear variational equation: Ha(uc, v) + gij'lu), v} = L(v),
VceF,
u,eV,
(6.27)
with
C Jfl^
Vw-Vv VwVv 2
- dx,
TvV| |V|2
V v, w e V.
(6.28)
If we define pt by p, = _
^
£
_
(6.29)
then PceA.
(6.30)
84
II Application of the Finite Element Method From (6.27)-(6.30) we obtain
lia(ue,v) + g | ps • Vv dx = L(v),
VveF, uceV.
(6.31)
Since A is a bounded closed convex subset of L2(Q) x L2(Q), it is weakly compact, so that from (p£)e we can extract a subsequence, still denoted by (p£)£, such that lim pc = p weakly in L2(O) x L2(£l),
p e A.
(6.32)
Actually we have pe -> p in L°°(n) x L°°(O) weakly *. Taking the limit as e -> 0 in (6.31), from (6.26) and (6.32) we have /M(U, v) + g \p-Vvdx
= L(v),
VtieF,
u e V,
(6.33)
so that (6.17) is proved. To complete the proof of the theorem, we need only prove that p-Vu = |V«|a.e..
(6.34)
Taking v = u in (6.33) and comparing with (6.22), we obtain
[\Vu\dxJJJ
[p-S/udx= JQ
[ (|Vu| -p-Vu)dx
= 0.
(6.35)
JQ
Since p e A, it follows from Schwarz inequality in U2 that p • Vu < | V« | a.e.. Combining (6.35) and this inequality we obtain p • Vu = |VM| a.e.. This proves (6.18) and, hence, the theorem.
•
Remark 6.3. The function p occurring in (6.17), (6.18) is not unique if Q e R 2 ; this is shown in G.L.T. [2, Chapter 5], [3, Chapter 5]. Remark 6.4. Relation (6.17) implies -nAu-gV-p=fmQ (6.36) u| r = 0. 6.7. Finite element approximation of (6.1) In this section we follow G.L.T. [3, Chapter 5]. For the sake of simplicity, we shall assume that Q is a polygonal domain of U2. 6.7.1. Definition of the approximate problem Let 9~h be as in Sec. 2 of this chapter. We approximate V by Vh = {vh e C°(Q), vk = 0onr,vh\TeP1,
V T G 3Th)
6. A Second Example of EVI of the Second Kind: Flow of Viscous Plastic Fluid in a Pipe
85
and (6.1) by Ha(uh, vh - uh) + gj(vh) - gj(uh) > , vh - uh),
V vheVh,
uheVh. (6.37)
Then: Theorem 6.4. The approximate problem (6.37) has a unique solution. Remark 6.5. So far, only an approximation by piecewise linear finite elements has been considered. This fact is justified by the existence of a regularity limitation for the solutions of (6.1), which implies that even with very smooth data we may have u £ H3(Q) (see Sec. 6.5). Nevertheless, in Fortin [1], Bristeau [1], G.L.T. [3, Chapter 5], and Bristeau and Glowinski [1], one may find applications of piecewise quadratic finite elements, straight or isoparametric, for solving (6.1). The numerical results which have been obtained seem to prove that for the same number of degrees of freedom the accuracies at the nodes are of the same order for the finite elements of order 1 and 2. From the above works it also appears that the second-order finite elements are much more costly to use (storage, computational time, etc.). We must also note that when using first-order finite elements, | n | Vvh | dx can be expressed exactly with respect to the values of vh at the nodes of 3Th, while with secondorder finite elements we need a numerical integration procedure. Remark 6.6. From the symmetry of a( •, •), (6.37) is equivalent to the minimization problem J(uh)<J(vh),
VvheVh,
uheVh,
(6.38)
where J(vh) = % a(vh, vh) + g f | \ v h \ dx - , vh}. z Jn
(6.39)
6.7.2. Convergence of the approximate solutions (general case) We use the notations of the previous sections. Theorem 6.5. / / , as h -> 0, the angles of STh are bounded from below uniformly in h, by 90 > 0, then lim \\uh- u\\v = 0, where u and uh are, respectively, the solutions of (6.1) and (6.37).
(6.40)
86
II Application of the Finite Element Method
PROOF. In order to prove (6.40), we use Theorem 6.3 of Chapter I. Here we have to verify that the following three properties hold: (i) There exist U <= V, U = V, and rh:U^Vh
such that limft^0 rhv = v strongly
in V, V v 6 U. (ii) If Vi, -> v weakly in V as h -> 0, then lim inf ji,(vi,) > j(v).
(iii) lim^o jh(rhv) = j(v), VveU. Verification of (i). We take U = @(Q). Then 0 = tfj(fi) = V. Define rhv by rhveVh,
V v e tf'(O) n C°(Q), (6-41)
Then since rftu is the "linear" interpolate of v on 9~h, it follows from Ciarlet [1], [2], and Strang and Fix [1] that under the above assumptions on STh, we have .»(fi)
(6.42)
Then from (6.42) we find that limfc_»0 rhv = v strongly in #J(fl), V » e U . Verification of (ii). Since7h(yft) = X^X V », £ F t , (ii) is trivially satisfied. Verification of (iii). Since jft(yft) = ;(;;,,), V yA £ 1^ and from the continuity of }{•) on V, (iii) is trivially satisfied. Hence from (i), (ii), and (iii), it follows that uh-*u
strongly in V.
•
6.7.3. Convergence of the approximate solutions ( / e L 2 (Q)) From the regularity theorem, Theorem 6.2 of this chapter, we have H lliP(n) H«ll
<
llll
if Q is convex and T is sufficiently smooth. This property still holds if Q is a convex polygonal set. In this section we will always assume that Q is a convex polygonal domain. We have the following: Theorem 6.6. With assumptions on 3~h as in Theorem 6.5, we have \\uh - u\\v = O(h112).
(6.44)
PROOF. From (6.1) and (6.37) it follows that fia{u, uh — u) + gj(uh) - gj(u) > (/, uh - u), Ha(uh, vh - uh) + gj(vh) - gj(uh) > ( / , vh - uh), V vh e Vh, which imply, by addition, - u,uh-
u) < na(uh - u, vh - u) + gj(vh) - gj(u) + fia(u, vh-u)~
( / , vh - u),
V»,6F».
(6.45)
Since j( •) is a norm on V, using the Schwarz inequality in V and (6.45), we obtain - I k - u\\l < - \\vh - u\\l + gj(vh -u)
+ /M(U, vh-u)-
( / , vh - u),
V vh e Vh. (6.46)
6. A Second Example of EVI of the Second Kind: Flow of Viscous Plastic Fluid in a Pipe
87
Since u e HQ(Q.) n H2(Q), we have a(u, v) = —I Auv dx,
VceF.
Hence, from (6.46) we have - I k - u\\v < - \\vh - u\\v + gj{vh - u) + 2.
Z
(-/iAw - f)(vh - u) dx,
V vh e Vh,
JJJ
so that - \\uh - u\\2v < - \\vh - u\\v + gj(vh - u) \\LHai)\\vh -
(6.47)
u\\LHa),
Since ||A«||L2(n) is a norm equivalent to the // 2 (fl)-norm over H2(Q) n //J(Q), it follows from (6.2), (6.43) and (6.47) that ^\\Uk -
U \ \ y < ^\\Vh
- U\\y
mM
,)
V « , E Fh.
(6.48)
Since F is Lipschitz continuous, we have (cf. Necas [1]) H2(Q) a C°(Q). The solution u of (6.1) 6 H2(ii) cz C°(Q); using the above inclusion, we can define rhu by rhueVh,
(rhu)(P) = v(P),
VPeS,.
From the above assumptions on ,Th, we have (cf. Ciarlet [1], [2] and Strang and Fix [1]) \\rhu - u\\v < y^Httll^n,, 2
\\rku - u\\LHn) < y2h \\u\\Hzin),
(6.49) (6.50)
where yt and y2 are constants independent of h and u. Taking vh = rhu in (6.48), it follows from (6.43), (6.49), and (6.50) that (6.51) with 70 = yo(Q) and M(Q) = meas(Q). Hence, from (6.51) we have I k - ti|| = O(h"2). This proves the theorem.
•
Remark 6.7. If Q is an open interval of R, i f / e L 2 (Q), and if piecewise linear elements are used, then an O(h) estimate can be obtained for the approximation error; the proof of the above estimate is very close to the proof of Theorem 3.4 of Sec. 3.7.1 and is given in detail in G.L.T. [3, Appendix 5].
88
II Application of the Finite Element Method
6.8. The case of a circular domain with / = constant In this section we consider a particular case of the general problem (6.1) by taking O = {x e IR2, jx\
+ x\ < R},
L(v) = C f vdx,
(6.52)
C > 0.
(6.53)
Jn
6.8.1. Exact solutions and regularity properties The solution of (6.1) corresponding to (6.52), (6.53) is given in Sec. 6.5.2 of this chapter. We recall that if g < CR/2, then ueVnW2'
a
(Q),
u$Vn
H3(Q).
(6.54)
In the sequel we assume that g < CR/2. 6.8.2. Approximation by finite element of order 1 Let &h be a finite triangulation of Q satisfying (2.22), (2.23) of Sec. 2.5, Chapter II and TCZQ.
(6.55)
Define Qh and Fh by
Then Qh <= Q, and in the sequel we assume that Th satisfies: all the vertices of Th belong to T.
(6.56)
Then we approximate V by Vh = {vh e C°(nh), vh = 0 on rh, vh\T ePltVTe
3Th}.
Now Vh can be considered to be a subspace of V, obtained by extending vh e Vh to Q by taking zero in Q — Q,,. It is then possible to approximate (6.1) by r
Ha{uh, vh - uh) + gj(vh) - gj(uh) > C
Jn
(vh - uh) dx,
V»,eFt,
uh e Vh. (6.57)
This is a finite-dimensional problem which has a unique solution. 6.8.3. Error estimate In this section we will obtain an error estimate of order hyj — log h. The following three lemmas play an important role in obtaining the above error estimate.
6. A Second Example of EVI of the Second Kind: Flow of Viscous Plastic Fluid in a Pipe
89
Lemma 6.1. Let p, q e IR2 - {0}. Then 2
IPl I q l
^
.
(6.58)
PROOF. We have
But
IPl
|q|2-2p-q = | p -
Iql
Consequently, j
^ <2|p-q|
IPl I q l
•
which obviously implies (6.58).
Remark 6.8. In (6.58), 2 is the best possible constant (take p = - q ) . Moreover, (6.58) is also true in UN, V N. Lemma 6.2. Let u and uh be the solutions o/(6.1) and (6.57), respectively. Let p satisfy (6.17), (6.18); then we have fia(uh - u,uhV vh e Vh,
u) < na{uh - u,vhand
V ph e A
u) + g
(ph - p) • V(vh - u) dx, •in
such that ph • Vvh = | Vvh | a.e..
(6.59)
PROOF. We shall prove (6.59) with feV*. From (6.45) we have Ha(uh -u,uh-u)<
iM(uh -u,vh-u)
+ gj(vh) - gj{u) + JM(U, vh - u)
Taking into account (6.17), we obtain lia(uh
- u , u
h
- u)< na(uh
- u, vh - u) - 9J((u) - g \ p-V{vh-u)dx, •In
(6.60)
Let vh e Vh and ph e A such that ph •Vvh = \ Vvh \ a.e.; such a ph always exists. Substituting this in (6.60) and using the relations •/(«») =
\ Ph • Vvh dx, •In
j{u)= I \Vu\dx> I ph-Vudx, we obtain (6.59). This proves the lemma.
(6.61) (6.62)
•
90
II Application of the Finite Element Method
Let u be the solution of (6.1) and 5 > 0. Define Q? c Q by Qs = {xeQ,
|Vw(x)| > 3}.
In the case of the problem (6.1) associated with (6.52), (6.53) (assuming g < CR/2), we have: Lemma 6.3. We have the following identity
(6.63) PROOF. From (6.16) we obtain du ~dr so that r dr
which implies (6.63).
D
From the above lemmas we deduce: Theorem 6.7. Let u be the solution of the problem (6.1) associated with (6.52), (6.53). Let uh be the solution of the problem (6.57) with S~h satisfying (6.55), (6.56). Assume that as h -> 0, the angles of3~h are bounded from below uniformly in hby 60 > 0. Then we have II"* — "IIK = OQiJ-logh).
(6.64)
PROOF. Starting from Lemma 6.2, from (6.59) we obtain ^.\\uh - u\\l < ~\\rhu - u\\l + g \ \ph2 2 Jn
p\\V{rhu - u)\dx,
V^eA
such that ph • Vrh u = | Vrh u \,
(6.65)
where rhu is defined by rhueVh,
(rhu)(P) = u(P),
V P e vertex of erh.
We have rhu = 0 on fi — Qh so that f — u\\y =
C \V(rhu — u)\2 dx =
C \Vu\2 dx +
\V(rhu — u)\2 dx. (6.66)
6. A Second Example of EVI of the Second Kind: Flow of Viscous Plastic Fluid in a Pipe
91
Let us define
\Vu\2dx, \V{rhu - u)\2 dx.
= ~{ 1
•'fih
It is easily shown that meas(Q - Qft) < - h2.
(6.67)
Furthermore, (6.16) implies |Vu(x)| < —\R - —I, 2/i \ CJ From (6.67) and (6.68) it follows that
Vx e a
^
2
2
.
(6.68)
(6.69)
Since u e W2< °°(Q), on each triangle T 6 f , , w e have (cf. Ciarlet and Wagshal [1]) IVfou - u)(x)\ < - 7 ^ - \\p(D2u)\\LX(n).
(6.70)
sin t/Q
where D2 u(x) is the Hessian matrix of u at x, defined by d2u
/ ___M
I D2u(x) =
8x\
3xj ox2
I 52u \ 3 5 ~ (x)
fe)x
52M
Tl. (x
yjXi dx2
8x2
and p{D2u{x)) is the spectral radius of D2u(x). We have 2o Z)2w(x) = 0 i f O < r < ~ ,
(6.71)
so that p(D2 u(x)) = 0, and it is easily verified that 7
C
9/7
2n
C
< R.
(6.72)
Then (6.70)-(6.72) imply (6.73) So it remains to estimate the termg \n\ph — p\\V(rhu — u)\ dx in (6.65). We have fi = U Q i;
(6.74)
92
II Application of the Finite Element Method
where
n 3 = a - nft, Q4 = {x e ilh, r>2g/C + h}, Q5 = {x e Qft, 2#/C - h
f
—J'
(6.75)
We have rhu = 0 over Q — Qfe, so that in (6.65) we can take ph = 0 o v e r n - Q , , .
(6.76)
From (6.67), (6.68), (6.70), and p e A, it follows that
I V« | dx < ^ ffcC« - ^ ) ^ 2 -
*3 < 0 f
(6-77)
From (6.70)-(6.72), and since p, ph e A, we have
so that X
5
< ^ ^ - ^ . /i sin 0O
(6.78)
From the definition of h(h = maximal length of the edges of T e f J , we have rhu = u = constant over Q6, so that X6 = 0.
(6.79)
It remains to estimate XA. Since the equipotential lines of u in fi4 are circular, for h sufficiently small, from (6.16) we have rhu\T # constant,
VTe^,,
T a Q4,
so that V^Mj-# 0,
VTe^,
T c H4.
(6.80)
Taking into account (6.80), it follows from (6.65) that Vrhu
VTe/,,
TcQ4.
(6.81)
Furthermore, we observe that (6.16) implies — >0, 2H
VxeQ4.
(6.82)
6. A Second Example of EVI of the Second Kind: Flow of Viscous Plastic Fluid in a Pipe 93 This, in turn, implies VxeO 4 . I Vu(x) I Applying Lemma 6.1 to the pair {Vrhu, Vu} and Lemma 6.3 with d = Ch/2[i, it follows from (6.80)-(6.82) that p
nu
Lwu ,|Vt +
where From (6.70) and (6.82) it follows that C 2 ( ChVc
dx
which, using (6.63), implies
or, more simply, XA<—
1 \R
ysin OQJ \
log — ]. C
(6.83)
RJ
Taking (6.65) into account, the estimate (6.64) is obtained by addition of the Xh i — 1,..., 6. More precisely, for sufficiently small h, Ik - u\\v < 4^ ^/n—- V-log *. sin 99 0 H sin
(6.84)
a 6.8.4. Generalization From the numerical experiment we have done, it seems that in a great number of cases (important from the point of view of application) we have the following properties for u: (1) we V n W2'X{£1), (2) Q o = {x | VM(X) = 0} is a compact subset of Q with a smooth boundary, (3) Q o has a finite number of connected components. Moreover, it seems that in the above cases we can conjecture that for S > 0, we still have
J n «|V«(x)| With these properties we can easily prove the following error estimate: \\uh-u\\v
= OQiJ - log h).
94
II Application of the Finite Element Method
Remark 6.9. Using an equivalent formulation of (6.1) (less suitable for computations) Falk and Mercier [1] have obtained an 0(h) estimate for ||«2 — u||jji(n) f° r a piecewise linear approximation (see also G.L.T. [3, Appendix 5]). 6.9. Iterative solution of the continuous and approximate problems by Uzawa's algorithm We begin with the continuous problem (6.1). Let us define if: V x H -> U by £C(v, q) = ~ a(v, v) - L(v) + g f q-Vvdx, *•
V»eF,
V qeH,
-In
where H = L2(Q) x L2(Q). Let {u, p} be the solution of (6.17), (6.18). Then we have: Theorem 6.8. The pair {«, p) is a saddle point of' $£ over V x A if and only if {u, p} satisfies (6.17) and (6.18). EXERCISE
6.5. Prove Theorem 6.8.
From Cea [2, Chapter 5] (see also G.L.T. [3, Chapter 5]) it follows that to solve (6.1) we can use the following Uzawa algorithm: p° e A arbitrarily chosen (for example, p° = 0):
(6.86)
then, by induction, knowing p", we compute u" and pn+1 by Ha(u", v) = < / » -
g
f / • Vv dx,
V v e V, u" e V,
(6.87)
Pn+1 = PA(p" + pgVW),
(6.88)
where PA: H —* A is the projection operator in the H-norm, defined by P (a) =
q
AW
Sup(l,|
V " ~ +d ' (6.89) «-| r = 0. We shall give a direct proof for the convergence of (6.86)-(6.88) based on Theorem 6.3 of Sec. 6.6.
Theorem 6.9. Let un be the solution of '(6.87). Then if
0 < p< ^ ,
(6.90)
6. A Second Example of EVI of the Second Kind: Flow of Viscous Plastic Fluid in a Pipe
95
we have lim | K - u\\v = 0,
(6.91)
n~* oo
where u is the solution of (6.1). PROOF. Let {u, p} satisfies (6.17) and (6.18). Then (6.18) implies V = PA(P + pgVu)-
(6-92)
We define u" = u" — u, p" = p" — p. Using the fact that P A is a contraction mapping, and from (6.88), (6.92), we obtain |p" + 1 1 2 < |p"| 2 + 2f>g f p" • Vu" dx + p2g2 I* |Vu"|2 dx, •'si
(6.93)
Jn
where
From (6.17) and (6.87) it follows that \xa(un, v) + g f p" • Vt; dx = 0, •in
V v e V.
(6.94)
Replacing v by «" in (6.94), we obtain \m(un, u") + g f p" • Vu" dx = 0.
(6.95)
From (6.93) and (6.95) we have |p"| 2 - | p " + T > p(2/i - pg2)\\u"\\2v.
(6.96)
2
If 0 < p < 2p/gf , then using standard reasoning, we obtain lim ||u"||F = 0, n-* ao
which proves the theorem.
•
Let us describe the adaptation of (6.86)-(6.88) to the approximate problem (6.37). W e define Lh c L 2 (Q) x L2(Q) by
where XT is the characteristic function of T. It is then clear that V « k e Fh, Vvh e Lh. We also define Ak by Aft = A n L,. We can easily prove that
Then (6.86)-(6.88) is approximated by: p° 6 Aft arbitrarily chosen,
(6-97)
96
II Application of the Finite Element Method
by induction, knowing pi, we obtain u\ and pl+1 by Ha(unh, vh) = L(vh) - g f p"h • Vvhdx,
VvheVh,
unheVh,
pl+1 = PA(pl + PgV<).
(6.98) (6.99)
2
Then for 0 < p < 2^/g we obtain the convergence of u\ to uh. EXERCISE
6.6. Study the convergence of (6.97)-(6.99).
Remark 6.10. The above methods have been numerically applied for solving (6.1) in Cea and Glowinski [2], Fortin [1], Bristeau [1], Bristeau and Glowinski [1], and G.L.T. [3, Chapter 5]. They appear to be very efficient and particularly well suited for taking nondifferentiable functionals like J n | Vv \ dx into account.
7. On Some Useful Formulae Let T be the triangle of Fig. 7.1. We denote by M(T) the measure of T. Let i; be a smooth function defined on T. We define v{ and vJk by vt = v(Mi),
vjk = v(Mjk).
Then we have the following formulae:
1
uv dx =
M(T)
+ v3)
12 V u, v e Pu
M«3 1 2 {\M2M3\2v\ vj 4M(T)2 + 2M2M\
+ 2M3M[
• M3M[V1V2
\\2v\ + \WjA*2\2v23 M2 •
+
M2M3v3v1 (7.2)
-MXM2V2V3},
M
M,
(7.1)
M2
Figure 7.1
7. On Some Useful Formulae
97
J \v\2 dx = —— {Joivi + x>\ + vj) + rs(v2l2 + v223 + v231) - v23v31 + v31v12) v3vl2)}, "' 2 dx
VveP2,
(7.3)
=
+ 2(t)12 + v23 - v31)M3M[\2 + \vxM2M\ + v2M3M[ - v3M1M2 + 2(v23 + v31 - v12)M^M*2\2 + | -vlM2M3 + v2M3M\ + v3MtM2 + 2(o31 + v12 -
v23)M2M'3\2}, \/veP2.
(7.4)
The above formula may be useful for expressing the approximations of the problems of this chapter, in a form suitable for computations.
CHAPTER III
On the Approximation of Parabolic Variational Inequalities
1. Introduction: References In this chapter we would like to give some indications of the approximation of parabolic variational inequalities (PVI) (mostly without proof). For a detailed reference, see Fortin [1], Bristeau [1], Bristeau and Glowinski [1], C. Johnson [1], and A. Berger [ I ] . 1 See also Richtmyer and Morton [1], Kreiss [1], and Lascaux [1] for the numerical analysis of time-dependent equations.
2. Formulation and Statement of the Main Results Let H and V be two real Hilbert spaces such that V c: H, V = H. Assuming that H = H*, we have V <= H c V*. The scalar product in H (resp., in V) and the corresponding norms are denoted by (•, •), H (resp., ((•, •)), II-ID- Moreover, we also use (•, •) for the duality between V* and V. We now introduce: • A time interval [0, T] with 0 < T < oo, a bilinear form a: V x V -> R, continuous and elliptical in the following sense: 3 a > 0 and X > 0 such that a(v, v) + /l|u| 2 > <x\\v\\2, V c e F , • fe L2(0, T; V*), u° e H (for the definition of L 2 (0, T; Z), see Lions [1],
[3]); • K: closed convex nonempty subset of V, • j:V -> U convex, proper, l.s.c. We then consider the following two families of PVI: Find u(t) such that tdu , v - u } + a(u, v -u)>(f,v -u), V v e K, a.e. t e ]0, T[, dt ) u(t) e K a.e. t e ]0, T[, w(0) = u0. 1
See also G.L.T. [3, Appendix 6] and the references therein.
(2.1)
3. Numerical Schemes for Parabolic Linear Equations
99
Find u(t) such that —, v - u I + a(u, v - u) + j(v) - j(u) >(f,v-
«), VveV, a.e. t e ]0, T[,
u(t) £ V a.e. t e ]0, T[, u(0) =
M0 .
(2.2)
Remark 2.1. If K = F a n d ; = 0, then (2.1) and (2.2) reduce to the standard parabolic variational equation: Y, v j + a(u, v) = (/, u),
M veV a.e. in t e ]0, T[,
«(t) e F a.e. t e ]0, T[,
M(0)
= u0.
(2.3)
Under appropriate assumptions on u 0 , K, and ;'(•), it is proved that (2.1), (2.2) have unique solutions in L 2 (0, T; V) n C°([0, T], if). For the proof we refer to Brezis [4], [5], Lions [1], and Duvaut and Lions [1]. In the following sections of this chapter, we would like to give some discretization schemes for (2.1) and (2.2), and then, in Sec. 6, study the asymptotic properties over time of a specific example, for the continuous and discrete cases.
3. Numerical Schemes for Parabolic Linear Equations Let us assume that V and H have been approximated (as h -* 0) by the same family (Vh)h of closed subspaces of V (in practice, Vh are finite dimensional). We also approximate (-,-)> a( •. •) by (•, • \ , ah{ •, •) in such a way that ellipticity, symmetry, etc. are preserved. We also assume that u0 is approximated by (uOh)h such that uOh e Vh and l i m ^ 0 uOh = u0 strongly in H. We now introduce a time step At; then, denoting u"h the approximation of u at time t = nAt (n = 0,1, 2,...), we approximate (2.3) using the classical step-by-step numerical schemes (i.e., we describe how to compute u"h+i if ui and u\~^ are known). 3.1. Explicit scheme (Euler's scheme)
n = 0, 1,... • u° = u
(3-1)
Stability: Conditional (See Lascaux [1] for the terminology). i>: O(At) (we consider only the influence of the time discretization).
100
III On the Approximation of Parabolic Variational Inequalities
3.2. Ordinary implicit scheme (backward Euler's scheme)
(<+1 ~< A +ah(uV\vh) = (fl+\vh)h, \ At 7»
\/vheVh,
n = 0,1,2,...;
u°h=uQh.
(3.2)
u°h = uOh.
(3.3)
Stability: Unconditional. Time accuracy: O(At).
3.3. Crank-Nicholson scheme 1
n+l n Uh — Uh
= 0,1,2,...;
Stability: Unconditional. Time accuracy: O(\At\2).
3.4. Two-step implicit scheme
= (/Z +1 ,«>*)*,
V^eF,,
n=l,2,...;
«? = u Oh , uj given.
(3.4)
Stability: Unconditional. Time accuracy: O(\At\2). Unlike the three previous schemes, this latter scheme requires the use of a starting procedure to obtain u\ from u° = uOh; to compute u\ we can use, for example, one of the schemes (3.1), (3.2), or (3.3); we recommend (3.3), since it is also an O(\ At|2)-scheme. Similarly, the generalizations of the scheme (3.4) discussed in Sees. 4 and 5 will require the use of a starting procedure which can be the corresponding generalization of schemes (3.1), (3.2), or (3.3).
3.5. Remarks Remark 3.1. The function /jj (or f"h+1/2) occurring in the right-hand sides of (3.1)-(3.4) is a convenient approximation of / at t = nAt (or t = (n + j)At). In some cases it may be defined as follows (we only consider / J since the technique described below is also applicable to f n h + 1 / 2 ) .
4. Approximation of PVI of the First Kind
101
First we define / " e V* by / " = / ( « At) if/eC°[0, T;V*l In the general case, it is defined by
!"+m)Atf(t)dt
f"=~ A r
ifn>l.
J(n-l/2)6a
Then, since (•, -)h is a scalar product on Vh, one may define f\ by (f"
u.V
= f f
v , )
V i). p V f
p
V
In some cases we have to use more sophisticated methods to define f"h. Remark 3.2. At each step (n + 1), we have to solve a linear system to compute u"h+1; however, if we can use a scalar product (•, • \ leading to a diagonal matrix, with regard to the variables defining vh, then the use of the explicit scheme will only require us to solve linear equations in one variable at each step. Remark 3.3. We can also use nonconstant time steps Atn. Remark 3.4. If we are interested in the numerical integration of stiff phenomenon or in long-range integration, we can briefly say: • Schemes (3.1), (3.2) are too dissipative; moreover the stability condition in (3.1) may be a serious drawback. • Scheme (3.3) is, in some sense, not sufficiently dissipative. • Scheme (3.4) avoids the above inconveniences and is highly recommended for "stiff" problems and long-range integration. In most cases the extra storage it requires is not a serious drawback. Remark 3.5. There are many works related to the numerical analysis of parabolic equations via finite differences in time and finite elements in space approximations. We refer to Raviart [1], [2], Crouzeix [1], Strang and Fix [1], Oden and Reddy [1, Chapter 9], and the bibliographies therein.
4. Approximation of PVI of the First Kind We assume that K in (2.1) has been approximated by (Kh)h, Kh <= Vh, V h, as in the elliptic case (see Chapter I). We also suppose that the bilinear form a(•, •) is possibly dependent on the time t and has been approximated by ah(t; uh, vh).
102
III On the Approximation of Parabolic Variational Inequalities
4.1. Explicit scheme + ah{n
\ >(flvh-u"h+\,
At;
u l Vh
-
u h+1)
"
u"h+leKh,
VvheKh,
n = 0,1,2,...;
u°h = uoh.
(4.1) Stability: Conditional (see G.L.T. [3, Chapter 6]). This scheme is almost never used in practice, since it is conditionally stable and the computation of u"h+i will generally, require the use of an iterative method, even if the matrix corresponding to (•, • )h is diagonal. 4.2. Ordinary implicit scheme ,n + 1 *
n *
' ,vh - unh+1 1 + ah((n + 1) A t ; unh+ \ vh -
u"h + leKh,
VvheKh,
u"h+1)
n = 0, 1, 2 , . . . ; «° = «ofc.
(4.2)
Stability: Unconditional. At each step we have to solve an EVI of the first kind related to Kh to compute u1+1. This scheme is often used in practice. 4.3. Crank-Nicholson scheme
+ a^n + *> At; "* Un 4-
u",+ll2sK, u
t
h
Jv
un+1
ul+112 /i'
u
h
—
«-012 j
'
"
—
u
>
1
'
z
• u° - u -' • • • '
u
h
—
"Oft-
(4.3)
Stability: Unconditional. Since (u",+ 1 - wJD/At = K + 1 / 2 - «J)/(Ar/2), we observe that at each step we have to solve an EVI of the first kind to compute unh+lj2. We also observe that possibly unh4 Kh. We do hot recommend this scheme if the regularity over time of the continuous solution is poor. 4.4. Two-step implicit scheme \ul+1 — 2ll"h + 2MJT1
At
n=
1,2,...;
u°h = uOh, u\ given.
(4.4)
5. Approximation of PVI of the Second Kind
103
Stability: Unconditional. At each step we have to solve an EVI of the first kind in Kh to compute u"k+1. Remark 3.4 also applies to this scheme.
5. Approximation of PVI of the Second Kind 5.1. Explicit scheme u"h+1) + jh(vh) - jh(u"h+1)
- ah(n At;u"h,vh>(flvh-u"h+\,
VvheVh,
ul+1eVh, n = 0,1,2,...;
u°h=uOh.
(5.1)
Stability: Conditional. This scheme also, is, almost never used in practice since it is conditionally stable and the computation of M£ + 1 will require the solution of an EVI of the second kind in Vh (in general, by an iterative method) even if the matrix corresponding to (•, •),, is diagonal. 5.2. Implicit scheme n+l _ *
>(fl
> Vh - K+1 ) + ah((n + 1) At; u\+ \ vh - unh+ J ) + jh(vh) +
- jh(u"h+ x )
+1
\vh-ul+\,
VvheVh, u"h"+1eVh, n = 0,1,2,...;
u°h=uOh.
(5.2)
Stability: Unconditional. At each step we have to solve in Vh an EVI of the second kind to compute
5.3. Crank-Nicholson scheme + ah((n + ^)At;utll2,vh ,vh-u"h+1'2\, U h+ h
" f
-;
VvheVh,
, n = 0,1,2,...;
Wh+l'2eVh, uho = uOh.
(5.3)
104
III On the Approximation of Parabolic Variational Inequalities
Stability: Conditional. Since (u"h+1 - u%)/At = (unh+112 - u"h)/(At/2), we observe that at each step we have to solve an EVI of the second kind to compute unh +1/2. If the regularity over time of the solution is poor, we do not recommend this scheme.
5.4. Two-step implicit scheme
2Uh 2 + 2Uh
-2
v
+ ah((n + 1) At; u\+ \ vh - u"h+1) > ( / J + 1 , vh - unh+1), VvheVh,
n = 0,1,2,...;
u°h = uOh, u\ given. (5.4)
We use one of the above schemes (5.1)—(5.3) to compute u\, starting from < = uOh. Stability: Unconditional. At each step we have to solve, in Vh, an EVI of the second kind to compute w£+1. Remark 3.4 applies to this scheme as well. 5.5. Comments The properties of stability and convergence of the various schemes of Sees. 4 and 5 are studied in the references given in Sec. 1. In some cases, error estimates also have been obtained. In Fortin [1] and G.L.T. [2, Chapter 6], [3, Chapter 6 and Appendix 6], applications to more complicated PVI than (2.1), (2.2) are also given. For the numerical analysis of hyperbolic variational inequalities, see G.L.T. [2, Chapter 6], [3, Chapter 6] and Tremolieres [1].
6. Application to a Specific Example: Time-Dependent Flow of a Bingham Fluid in a Cylindrical Pipe Following Glowinski [4], we consider the time-dependent problem associated with the EVI of Chapter II, Sec. 6 and study its asymptotic properties.
6.1. Formulation of the problem. Existence and uniqueness theorem Let Q be a bounded domain of U2 with a smooth boundary F. We consider: • V = ffj(fi), H = L2(Q), V* = H-^Q), • a(u, v) = j n VM • Vv dx, • A time interval [0, T], 0 < T < oo,
6. Time-Dependent Flow of a Bingham Fluid in a Cylindrical Pipe
105
. / G L 2 ( 0 , T ; V*),u0eH, •700 = $n\Vv\dx, • n > 0, g > 0. We then have the following: Theorem 6.1. The PVI (du
,v - u\ + \ia(u, v - u) + gj(v) - gj(u) >(f,v-
u),
V v e V a.e. t e ]0, T[, u(x, 0) = uo(x), (6.1) has a unique solution u such that u E L2(0, T;V)n
C°([0, T ] ; if),
- ^ e L 2 (0, T, F*)
a«d tfcjs V u0 6 # , V / e L2(0, T; F*). For a proof, see Lions and Duvaut [1, Chapter 6].
6.2. The asymptotic behavior of the continuous solution Assume that / is independent of t and that / € L2(Q). We consider the following stationary problem: Ha(u, v - u) + gj(v) - gj(u) > (/, v - u),
V v e V, u e V.
(6.2)
In Lions and Duvaut [1, Chapter 6] (see also Chapter II, Sec. 6 of this book), it is proved that u = 0 if#j3> ||/|| L 2 ( O ) ,
(6.3)
B = inf . v,v Mmm
, , .. (6-4)
where
t>#0
Then we can prove the following: Theorem 6.2. Assume that f e L2(Q) with ||/|| L 2 ( n ) < fig; then if u is the solution of (6.1), we have u(t) = 0 for t > A
oA*
\
py -
117 IILV (6-5)
where Ao is the smallest eigenvalue of — A in H{,(£1) (2.0 > 0).
106
III On the Approximation of Parabolic Variational Inequalities
PROOF. We use | -1 for the L2(fi)-norm and || • || for the Ho(fl)-norm. Since
it follows from Theorem 6.1 that the solution of (6.1) is defined on the whole IR + . We now observe that if gfl > | / 1 , then zero is the unique solution of (6.2); it then follows from Theorem 6.1 that if u(t0) = 0 for some 10 > 0, then u(f) = 0,
Vt>f0.
(6.6)
Taking v = 0 and v = 2w in (6.1), we obtain '8u \ —, u I + fia(u, u) + gj(u) = ( / , u) a.e. in f.
(6.7)
But since u e L 2 (0, T; V), »' e L 2 (0, T ; V*) implies (this is a general result) that t -> |u(t)l 2 is absolutely continuous with (d/dt)\v\2 = 2(dv/dt, v), from (6.7) we obtain - - | u | 2 + ixa(u, u) + gj(u) = (/, u) < \ f \ \ u \ a.e. in t. Since a(v, v) > ko\v\2, obtain
(6.8)
V c e V, and j(v) > f}\v\, V n e F (from (6.4)), from (6.8) we
|u| 2 + H ) M 2 + (gP - \f\)\u\
< 0 a.e. in teU+.
(6.9)
2 dt Assume that u(t) # 0, V t > 0; since t -> | u(i) \2 is absolutely continuous with | u(t) | > 0, it follows that t -> | u(t) | is also absolutely continuous. Therefore from (6.9) we obtain - I / I ) < 0a.e. te R + .
(6.10)
< -nk0
(6.11)
From (6.10) it follows that
d/dt\u(t)\
a.e. t e R+.
Define y by y = (g/S — \f\ )//U 0 ; then y > 0. By integrating (6.11) it then follows that |«(f)l + y < ( | « 0 | + y)e~"Xo\
Vt6R+;
(6.12)
(6.12) is absurd for t large enough. Actually we have u(t) = 0 if
i.e., (6.13)
6.1. Let fe L2(Q), possibly with | / 1 > gfi. Let us denote by u^ the solution of (6.2); then prove that EXERCISE
where u(t) is the solution of (6.1).
6. Time-Dependent Flow of a Bingham Fluid in a Cylindrical Pipe
6.3.
107
On the asymptotic behavior of the discrete solution
We still assume that / e L2(fi). To approximate (6.1) we proceed as follows: Assuming that Q is a polygonal domain, we use the same approximation with regard to the space variables as in Chapter II, Sec. 6 (i.e., by means of piecewise linear finite elements, see Chapter II, Sec. 6). Hence we have ah(uh, vh) = a(uh, vh),
\/uh,vheVh,
and from the formula of Chapter II, Sec. 7, we can also take («*,
vh)h = (uh, vh),
M
uh,vheVh.
Then we approximate (6.1) by the implicit scheme (5.2) and obtain n+l
.n
gj(vh)
At -gj(wh+1)>(fh,vh-unh+1), u"h+leVh;
VvheVh,
n = 0,1,2,...;
u°h = uOh.
(6.14)
We assume that uOheVh,^ h and lim uOh = M0 strongly in L2(Q).
(6.15)
Similarly, we assume that / is approximated by (fh)h in such a way that (/*» vh) can be computed easily and lim fh = / strongly in L2(Q). Theorem 6.3. Let \f\ < fig. If (6.15) and (6.16) /JOW, small, we have wj| = 0 for n large enough.
(6.16) then if h is sufficiently
PROOF. AS in the proof of Theorem 6.2, taking vh = 0 and u,, = 2i(j + : in (6.14), we obtain \-^—.
-, u"h+1)
+ / M \ V u " h + l \ 2 dx + g
| V K £ + 1 | dx =
fhu"h+1
dx,
Vn> 0; (6.17)
2
using the Schwarz inequality in L (Q), it follows from (6.17) that At——
|Mft+1l + M ) K + T + {gP - l/J)l"i!+1l ^ o,
Vn > o. (6.18)
2
Since fh-* f strongly in L (il), we have gP ~ I/*I > 0 for h sufficiently small
(6.19)
108
III On the Approximation of Parabolic Variationai Inequalities
From (6.18), (6.19) it then follows that u"h° = 0 => u"h = 0 for n > n0 if h is small enough.
(6.20)
Assume that uj ^ 0, V n; then (6.18) implies V»>0.
(6.21)
We define •/,, by yh = g(3 - \ fh |; then yh > 0
for h small enough and
lim yh — y = g[l — \ f |.
(6.22)
/i->0
From (6.21) it follows that
which implies that Vi
Since yh > 0 for h small enough, (6.23) is impossible for n large enough. More precisely, we shall have u\ = 0 if I'll
^
/1
which implies: vl
/ / h is small enough, then u"h = 0 if n >
.
(6.24)
Log(l + lon At) Relation (6.24) makes the statement of Theorem 6.3 more precise. Moreover, in terms of time, (6.24) implies that u"h is equal to zero if (6.25)
Log(l + Xoii At) We observe that Log[l + Xoii{\u°\lyJ\
1
/
|«°|
hm Af — - — — — = - — Log 1 + X0fi Log(l +XolxAt) Aoyu \ y h^0
Hence, taking the limit in (6.25), we obtain another proof (assuming that unh converges to u in some topology) of the estimate (6.5) given in the statement of Theorem 6.2. EXERCISE 6.2. Let u£° be the solution of the time-independent problem associated with/,,, possibly with \fh\ > fig; then prove that
Arr" / 2 K - K\,
V« > 0.
6. Time-Dependent Flow of a Bingham Fluid in a Cylindrical Pipe
109
6.4. Remarks Remark 6.1. We can generalize Theorem 6.2 to the case of a Bingham flow in a two-dimensional bounded cavity. Remark 6.2. In Glowinski [4], Bristeau [1], and Begis [1], numerical verifications of the above asymptotic properties have been performed and found to be consistent with the theoretical predictions. Remark 6.3. In H. Brezis [5], one may find many results on the asymptotic behavior of various PVI as t -> + oo.
CHAPTER IV
Applications of Elliptic Variational Inequality Methods to the Solution of Some Nonlinear Elliptic Equations
1. Introduction For solving some nonlinear elliptic equations, it may be convenient, from the theoretical and numerical points of view, to view them as EVFs. In this chapter we shall consider two examples of such situations: (1) a family of mildly nonlinear elliptic equations; (2) a nonlinear elliptic equation modeling the subsonic flow of an inviscid compressible fluid.
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
2.1. Formulation of the continuous problem Let Q be a bounded domain of UN (N > 2) with a smooth boundary T. We consider . V = tf i(Q); • a: V x V -> U bilinear, continuous, and F-elliptic with a > 0as ellipticity constant; a(-, •) is possibly nonsymmetric; • <j>: U. -> U, 4> e C°(R), nondecreasing with 0(0) = 0. We then consider the following nonlinear elliptic equation (P) defined by: Find u e V such that a(u, v) + <>(«), v} = L(v),
VceK,
From the Riesz Representation Theorem it follows that there exists A e Se(V, V*)
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
111
such that a{u, v) = {Au, v}, V « , c e F . Therefore (P) is equivalent to Au + (j)(u) =f,
ueV,
(2.1)
X
EXAMPLE 2.1. Let us consider a function a0 e L (Q) such that ao(x) > a. > 0 a.e. in Q.
(2.2)
a(u, v) = | ao(x)Vw • Vu dx + ) j8 • VM t?dx,
(2.3)
Define a(-, •) by
where y? is a constant vector in UN. From the definition of ao(-\ and using the fact that j n )S • Vv vdx = 0, V v e HQ(Q), we clearly have a(»,tO 2: a||t>||£.
(2.4)
Au = -V-(a 0 Vw) + p- Vu.
(2.5)
From (2.3) we obtain
Hence, in this particular case, (2.1) becomes - V • (a 0 Vu) + P • Vu + (j>(u) = / ,
u G F, (/.(«) e L J (Q).
(2.6)
Remark 2.1. IfN = 1, we have Ho(fi) <= C°(O). Because of this inclusion there is no great difficulty in the study of one-dimensional problems of type (P). If N > 2, the main difficulty is precisely related to the fact that ffo(fi) is not contained in C°(Q). Remark 2.2. The analysis given below may be extended to problems in which either V = H1(Q) or V is a convenient closed subspace of H^Q).
2.2. A variational inequality related to (P) 2.2.1. Definition of the variational inequality Let
=f
(2.7)
Jo
(2.8) The functional;: L 2 (Q) -> U is defined by j(v) = f O(D) rfx if
j(f) = + oo if ®(v) $ L\Q).
(2.9)
112
IV Applications of Elliptic Variational Inequality Methods
Instead of studying the problem (P) directly, it is natural to associate with (P) the following EVI of the second kind: a(u, v — u) + j(v) — j(u) > L(v — u),
V v e V, ueV.
(n)
If a(-, •) is symmetrical, a standard method for studying (P) is to consider it to be the formal Euler equation of the following minimization problem encountered in the calculus of variations: J(u) < J(v),
V v e V, ueV,
(2.10)
where J(v) = %a(v, v) + I 3>(u) dx - L(v). EXERCISE
(2.11)
2.1. Prove that D(
2.2.2. Properties of ){•) Since >: U -» U is nondecreasing and continuous with (/>(0) = 0, we have $eC'(R),
O(t) > 0, V t e R .
(2.12)
The properties ofXO are given by the following: Lemma 2.1. The functional j(-) is convex, proper, and l.s.c. over L2(Q). PROOF. Since j(v) > 0, V v e L2(£l), it follows that ;(•) is proper. The convexity of ;(•) is obvious from the fact that O is convex. Let us prove that;'(•) is l.s.c. Let {vn}n, vn e L2(Ji), be such that lim i!n = v strongly in L2(Q). Then we have to prove that lim inf ;(*;„) >7(i>).
(2.13)
If lim inf,,^ j(vn) = + oo, the property is proved. Therefore assume that lim inf j(vn) = I < +co. Hence we can extract a subsequence {vnu}nk such that ) = /,
(2.14)
vnk -* v a.e. in Cl.
(2.15)
lim O(iO =
(2-16)
l
Since O e C (R), (2.15) implies
Moreover,
(2.17)
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
113
Hence, by Fatou's Lemma, from (2.16) and (2.17), we have
lim inf | ®(vnj dx > ) O(u) dx.
(2.18)
From (2.14) and (2.18) we obtain (2.13).
This proves the lemma.
•
Corollary 2.1. The functional j(-) restricted to V is convex, proper, and l.s.c.. 2.2.3. Existence and uniqueness results for (n) Theorem 2.1. Under the above hypotheses on V, a(-, •), L(-), and >(•), problem (n) has a unique solution in V n PROOF. Since V, a(-, •), L(-), and j(-) have the properties (cf. Corollary 2.1) required to apply Theorem 4.1 of Chapter I, Sec. 4, the EVI of the second kind, (n), has a unique solution u in V. Let us show that u 6 D(9). Taking v = 0 in (n), we obtain a(u,u) +j(u) < L(u) <\\f\\Ju\\v.
(2.19)
Since j(u) > 0, using the ellipticity of a{-, •), we obtain Nlv£—. a
(2-20)
j(u)<^-.
(2.21)
which implies
a This implies u e D(
•
Remark 2.3. If a(-, •) is symmetric, (TT) is equivalent to (2.10).
2.3. Equivalence between (P) and in) In this section we shall prove that (P) and (n) are equivalent. First we prove that the unique solution of (n) is also a solution of (P). In order to prove this result, we need to prove that >(M) and u(j>(u) belong to L^Q). Proposition 2.1. Let u be the solution of (n). Then ucj)(u) and (j)(u) belong to L\Q). PROOF. Here we use a truncation technique. Let n be a positive integer. Define
Kn = {veV,\v(x)\
114
IV Applications of Elliptic Variational Inequality Methods
Since Kn is a closed convex nonempty subset of V, the variational inequality a(un, v - un) + j(v) - j(un) > L(v - Mn),
V v e Kn,
uneKn,
(nn)
has a unique solution (in order to apply Theorem 4.1 of Chapter I, we need to replace j byj + lKn, where IKn is the indicator functional of Kn). Now we prove that lim,,.,^ un = u weakly in V, where u is the solution of (n). Since Qe Kn, taking v — 0 in (nn), we obtain (as in Theorem 2.1 of this chapter) \\un\\v <
=,
(2.22)
ex (2.23)
From (2.22) it follows that there exists a subsequence {unk}nk of {«„}„ and u* e V such that lim unH = u* weakly in V.
(2.24)
k->oo
Moreover, from the compactness of the canonical injection from HQ(Q) to L2(Q) and from (2.24), it follows that lim unk = u* strongly in L2(Q).
(2.25)
Relation (2.25) implies that we can extract a subsequence, still denoted by (unjnk, such that lim unk = u* a.e. in il.
(2.26)
Now let v e V n L ^ f l ) ; then for large k, we have v e Knk and «(«»„> "»„) + JKJ < a{unk, v) + j(v) - L{v - uj.
(2.27)
Since lim infk^^ a(unk,unk) > a(u*,u*) and lim inf^..^ _/(wBk) > j(w*), it follows from (2.24) and (2.27) that a(u*, «*) + j(«*) < a(u*, v) + j(v) - L(v - «*),
V t; e L ^ f i ) n K,
M* 6 V,
which can also be written as
a(u*,v-u*)+j(v)-j(u*)>L(v-u*),
V v e V n L°°(ii), «* e K (2.28)
For n > 0, define rn: F -» Kn by rnv = Inf(n, Sup(-n, y))
(2.29)
(see Fig. 2.1). Then, from Corollary 2.1 of Chapter II, Sec. 2.2, we have lim tn v = v strongly in V, (2.30) lim xnv = v a.e. in Q. n-* a:
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
115
Figure 2.1 Moreover, we obviously have Tnu(x)| < |u(x)|a.e.,
(2.31)
v(x)inv(x)>0a.e..
(2.32)
From (2.30)-(2.32) and from the various properties of
(2.33)
limO>(Tny) = ®(v)a.e..
(2.34)
n-» oo
Since xnv e L°°(n) n V, it follows from (2.28) that a(u*, xnv-u*)+
j(xnv) - j(u*) > L(xnv - u*\
V ve V, u* e V. (2.35)
If v $ D(Q>), then by Fatou's lemma, \imj'(Tnv) = +oo. n-* co
If v e D(
From these convergence properties and from (2.30), and by taking the limit in (2.35), it follows that a(u*, v -u*)+ ){v) - j(u*) > L(v - u*\
VveV,
u* e V.
(2.36)
Then u* is a solution of (n), and from the uniqueness property, we have u* = u. This proves that lim,,^ un = u weakly in V. Let us show that <j>(u), u
dx>L(v-un),
VveKn. (2.37)
116
IV Applications of Elliptic Variational Inequality Methods
Since 0> e C\U) and
* ( « „ + t(p - «„)) - O(u B )
= <£(«„)(» - un) a.e..
(2.38)
Moreover, since <J> is convex, we also have, V t e ]0, 1], «0O0
n)
*
From (2.38), (2.39), and using Lebesgue's dominated convergence theorem in (2.37), we obtain a(un,v-un)+
\
VceX,.
(2.40)
Then, taking v = 0 in (2.40), we have a(un, «„) +
4>{un)un dx < L(un),
which, using (2.22), implies II f\\2
C
UkK
Jn
(2.41)
a
But 4>{v)v > 0, V v e V. Hence 4>(un)un is bounded in L1 (Q). Moreover, for some subsequence {«»Jnt of {«„}„, we have >(«„>«„ - * »(")"
a
-
e
-
i n
a
X
Then by Fatou's lemma we obtain u<j>{u) e L {Q), and this completes the proof of the proposition, since u(j>(u) e L J (Q) obviously implies that 4>(u) e L^Q). D Incidentally, when proving the convergence of (un)n to u, we have proved the following useful lemma: Lemma 2.2. The solution u of (n) is characterized by a(u, v — u) + j(v) — j(u) > L(v — w), VueFnL^Q),
ueV, <&{u)eL\Q).
(2.42)
In view of proving that (71) implies (P), we also need the following two lemmas: Lemma 2.3. The solution u of (n) is characterized by a(u, v — u) +
(j>(u)(v — u) dx > L(v — u), V i ) € L°°(fi) n V, ueV, u<j>(u) e L\Q).
(2.43)
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
117
PROOF. (1) (7c) implies (2.43). Let v e L°°(f2) n V. Then v e £>(), and since D(), V t e ]0, 1]. Replacing v by u + t(v — u) in (7c) and dividing by t, we obtain V t e ]0, 1], a(u,v-u)+
\ (2-44)
Since O e C 1 and is convex, we have <Wu + t(v — M)) — <1>(M)
lim —
>1
\1
=
^(U)(,, _ M) a .e.,
(2.45)
f
t->o 00
^HXo _ H) < ^ + ^ - ».)) -
(2 46)
By Proposition 2.1, we have <^(M), u(j>(u)e L\Q). Hence ^(u)(f - tOeL^fi) and (u) — O(M) e L'(£2), Vu e Lcc(f2) n K Then, using the Lebesgue dominated convergence theorem, it follows from (2.45) and (2.46) that r O(u + t(y - «)) - 4>(M)
f
Using the above relation and (2.44), we obtain (2.43). This proves that (7c) => (2.43). (2) We will now prove that (2.43) => (n). Let u be a solution of (2.43). Since $ is convex, it follows that - O ( M ) = O(0) - 0>(M) > <j){u)(O - u) = -0(«)u. This implies 0 < 0>(w) < «>(«) and $(u) e L^fl). Let w 6 L^fl) n K Then, from the inequality 4>(u)(v — u) <
j(v) - j(u),
which combined with (2.43) and
V v e L°°(Q) n F,
« e V,
by <j)(u) e
L^O). (2.47)
PROOF. (1) (TC) implies (2.47). Let v e V n L°°(fi). If u is the solution of (71), then u is also the unique solution of (2.43).
118
IV Applications of Elliptic Variational Inequality Methods
Let xn be denned by (2.29). Then xnu e V n Lm(Q). Replacing v by xnu + v in (2.43), we obtain a(u, v) +
4>(u)v dx + a(u, xnu — u) + J JJ
(f>(u)(xnu — u) dx > L(v) + L(xnu — u),
JQ
V c e F n L°°(n).
(2.48)
From (2.29)-(2.32) it follows that lim a(u, xnu — u) = 0, (2.49) lim L(xnu — u) — 0, lim 4>(u)(xnu -u)
= Q a.e.,
(2.50)
0 < 4>(u)(u - xnu) < u(j)(u) a.e..
(2.51)
Then, by the Lebesgue dominated convergence theorem and (2.50), (2.51), we obtain lim 4>(u)(xnu - u) = 0 strongly in L^fi).
(2.52)
n-* ao
Then (2.48), (2.49), and (2.52) imply a(u, v) + I 4>(u)v dx > L(v),
V t e F n LCO(O).
Since the above relation also holds for — v, we have a(u, v) + I* 4>(u)v dx = L(v),
V v e V n L°°(Q).
(2.53)
By Proposition 2.1 we have 4>{u) e L^fi); combining this with (2.53) we obtain (2.47). This proves that (TC) => (2.47). (2) (2.47) implies (n). We have a(u, v) + J <j)(u)v dx = L(v),
Vv e V n
Lx(il).
Then a(u, xnu)+
I 4>{u)xnu dx = L(xnu),
V n.
(2.54)
Since Tn« -> u strongly in V, {|n cf>(u)xnu dx]n is bounded. But
(2.55)
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
119
Hence, by the Lebesgue dominated convergence theorem, lim
4>(u)Tnu dx =
4>(u)u dx,
which along with (2.54) yields a(u, u) + f (p(u)u dx = L(u).
(2.56)
•In
Then by substracting (2.56) from (2.47) we obtain a(u, v — u) +
<j>iu){v — u) dx = L(v — u), V c e F n L°°(fi),
u e V, u(j)(u) e L^fi), (2.57)
and obviously (2.57) implies (2.43). This completes the proof of the lemma.
•
Corollary 2.2. / / u is the solution of (n), then u is also a solution of (P). PROOF. We recall that V* = H ^ ^ Q ) c S>'(°-) and that a(u, v) = (Au, v), V u, v e V and L(v) = , v}. Let u be a solution of (n). Then u is characterized by (2.47), and since S(£l) <= V, we obtain
f
-'a -'a
t>>,
v e ®(fi).
(2.58)
From (2.58) it follows that Au + 0(«) = / i n @'(O.);
(2.59)
since Au a n d / e F*, we have 0(u) e F*. Hence
= V, there exists a sequence {vn}n,vn e ®(Q),
lim vn = v strongly in V.
(2.60)
n-* oo
Let us define wn by (see Fig. 2.2) wn = min(u+, yn+) - min(«", v~).
(2.61)
120
IV Applications of Elliptic Variational Inequality Methods
Figure 2.2 Then wn has a compact support in SI,
(2.62) (2.63)
IKIIt-tffl ^ Mli/»(n>> and it follows from Chapter II, Corollary 2.1 that lim wn = v strongly in V.
(2.64)
From (2.63) and (2.64) we obtain lim,,^ vvn = v for the weak* topology of L°°(fl). Thus we have proved that * = ( c e L°°(fi) n V, v has compact support in Q} m
is dense in L (fi) n V for the topology given in the statement of the lemma. Let ve"U and (pn)n be a mollifying sequence (see Chapter II, Lemma 2.4). Define vn by v(x)
if x e fi,
(2.65) (2.66)
Vn = Pn * V.
Then vn e 3>(UN),
lim SB = v strongly in i/1([RN),
Dn has a compact support in Q, for n large enough.
(2.67) (2.68)
Let vn = vn\n; then for n large enough, ;>„ e <3(Q), and lim,,-,^ vn = v strongly in V. ||L=o(n), it follows from (2.66) that Since IICIIi^duK) •)n(x)\ < \ pn(x-
y)\v(y)\dy<
\\v\\
(2.69)
From this it follows that
< INI
(2.70)
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
121
Summarizing the above information, we have proved that V t e L™(£2) n V, there exists a sequence (vn\, vn e @{Q), V n, such that lim vn = v strongly in V, K I I L - W , ^ Nli-
Vn.
(2.71) (2.72)
Hence from (2.71) and (2.72) we find that vn -> v in L°°(Q) weak *. This completes the proof of the lemma. • Theorem 2.2. Under the above hypothesis on V, a(-, •), L(-), arcd $(•)» problems (n) and (P) are equivalent. PROOF. We have already proved that (n) implies (P). We need only prove that (P) implies (n). From the definition of (P) we have a(u, v) + <0(tO, »> = L(v),
ViieF,
u e V, <j>{u) 6fl"'(Q) n L^fl).
(2.73)
From (2.73) it follows that . a(u, v ) +
f (t>(u)v dx = L ( D ) ,
V c e ®(0).
(2.74)
If u e F n Lco(f2), from Lemma 2.5 we know that there exists a sequence (vn\, vn e such that lim vn = v strongly in V,
(2.75)
n-*oo
lim vn = v in L*(n) weak *.
(2.76)
n-* cc
Since vn e 3i(Q), from (2.74) we have a{u,vn)+ U(u)vndx = L(vn).
(2.77)
From (2.77) it follows that lim,,^ a(u, vn) = a(u, v), limn^^ L(vn) = L(v), and since 4>{u) e L 1 ^ ) , (2.76) implies lim
4>(u)vn dx =
(j>(u)v dx.
Thus, taking the limit in (2.77), we obtain a(u, v) + f 4>(u)v dx = L(y),
V t e F n L°°(Q).
Therefore (P) implies (2.47) which in turn implies (;t). This completes the proof of the theorem. • EXERCISE 2.2. In IR2, find a function v such that v e H~ 1(Q) n L ^ Q ) , v $ L"(Q), V p > 1, where Q is some bounded open set in IR2.
122
IV Applications of Elliptic Variational Inequality Methods
EXERCISE 2.3. Prove that if u > 0 a.e., then
2.4. Some comments on the continuous problem We have studied (P) and (n) with rather weak hypotheses, namely, <j) e C°(U) and nondecreasing, and / e V*. The proof we have given for the equivalence between (P) and (7r) can be shortened by using more sophisticated tools of convex analysis and the theory of monotone operators (see Lions [1] and the bibliography therein). However, our proof is very elementary, and some of the lemmas we have obtained will be useful in the numerical analysis of the problem (P). Regularity results for problems a little more complicated than (P) and (TT) are given in Brezis, Crandall, and Pazy [1]; in particular, for 2 / G L 2 ( Q ) and with convenient smoothness assumptions on A, the H (Q)regularity of u is proved.
2.5. Finite element approximation of (n) and (P) 2.5.1. Definition of the approximate problem Let O be a bounded polygonal domain of U2 and let &], be a triangulation of Q. satisfying (2.21)-(2.23) of Chapter II. We approximate V by Vh = {vh e C°(Q), vh\r = 0, vh\T e i \ ,
VTefJ.
Then it is natural to approximate (P) and (n), respectively, by
a(uh,vh)+
I 4>{vh)vh dx = L(vh),
V»,eF», uheVh
(PJ)
and a(uh, vh - uh) + j(vh) - j(uh) > L(vh - uh),
V vheVh,
uheVh
{nf)
with j(vh) =
®(vh) dx.
Obviously (P%) and (n%) are equivalent. From a computational point of view, we cannot generally use (Pf) and (n%) directly, since they involve the computation of integrals which cannot be done exactly. For this reason we shall have to modify (n*) and (P*) by using some numerical integration procedures. Actually we shall have to approximate a(-, •), L(-), andj(0- Since the approximation of (•, •) and L(-) is studied in Ciarlet [1, Chapter 8], we shall assume that we still work with a(-, •) and L(-), but we shall approximate j(-).
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
123
Figure 2.3
M17
To approximate X 0 w e shall use the two-dimensional trapezoidal method. Hence, using the notation of Fig. 2.3, we approximate j ( 0 by
JM=
I
" " ^ ^ Z*(»»(Mir)),
VvheVh.
(2.78)
Actually 7fc(uA) may be viewed as the exact integral of some piecewise constant function. Using the notation of Chapter II, Sec. 2.5, assume that the set I.h of the nodes of 3~h has been ordered by i = 1, 2 , . . . , Nh, where Nh = Card(Lh). Let Mt e 2,h. We define a domain Q ; by joining, as in Fig. 2.4, the centroids of the triangles, admitting Mt as a common vertex, to the midpoint of the edges admitting Mt as a common extremity (if M ; is a boundary point, the modification of Fig. 2.4 is trivial). Let us define the space of piecewise constant functions
Lk = - L l ^ = t n,xt, /<« e R, i = 1, 2,..., Nh\
(2.79)
where %t is the characteristic function of Q ; . We then define qh: C°(Q) n #J(Q) -> Lft by
9»i; = £ v(Mt)Xi.
(2.80)
i=i
Then it follows from (2.79) and (2.80) that =
\
h)
dx,
Figure 2.4
V^K,
(2.81)
124
IV Applications of Elliptic Variational Inequality Methods
implying that MVH) = j(qhvh),
VvheVh.
(2.82)
Then we approximate (P) and (n) by a{uh, vh) +
(t>(qhuh)qhvh dx = L(vh),
V vh e Vh, uheVh
(Pft)
Jn and a(uh, vh - uh) + jh(vh) - jh(uh) > L(vh - uh),
V vheVh,
uheVh.
(nh)
Then: Theorem 2.3. Problems (Pft) and (nh) are equivalent and have a unique solution. EXERCISE
2.4. Prove Theorem 2.3.
2.5.2. Convergence of the approximate solutions Theorem 2.4. If, as h-> 0, the angles of ^h are uniformly bounded below by 90 > 0, then lim \\uh — u\\r = 0, where u and uh are, respectively, the solution of(P) and (Ph). PROOF. Since j(-) is not continuous on V, the result of Chapter I, Sec. 6 on the approximation of the EVI of the second kind cannot be applied directly. However, the proof of convergence follows the same lines as in Theorem 6.3 of Chapter I. (1) A priori estimates for uh. Taking vh = 0 in (;tft), we obtain (2.83) 0<
J
®(qhuh)dx<^l*.
Jn
(2.84)
a
(2) Weak convergence ofuh. From (2.83) and from the compactness of the injection of V in L2(Q), it follows that we can extract from (uh)h a subsequence, still denoted by (uh)h, such that uh -> u* weakly in V, (2.85) uh -> u* strongly in L2(Q), uh -* u* a.e. in
(2.86) ft.
(2.87)
Admitting, for the moment, the following inequality: Uh% - VHWLHSI) ^ -j IIVf ll^o) x u.(n).
VcjePi,
V p with 1 < p < oo, (2.88)
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
125
it follows from (2.83) and (2.86) that strongly in L2(Q).
qhuh^u*
(2.89)
Then, modulo another extraction of a subsequence, we have qhuh-> u* a.e. mil,
(2.90)
from which it follows that d>((jh uh) -»
(2.91)
Then taking v e S>(£2), it follows from Ciarlet [1], [2] and Strang and Fix [1] that under the assumptions on 2Th of the statement of the theorem, we have ^
)
(
(2.92)
)
2
IIrhv - v||L.(n) < Ch 1|v||„,,„<„,,
V v e @(Q),
(2.93)
where C is a constant independent oft) and h and rh is the usual linear interpolation operator over &~h. Moreover, (2.88) with p = + oo, (2.92), and (2.93) imply hm \\qhrhv - v\\L.m = 0,
V » e S>(fi).
(2.94)
Taking yh = rhv in (7th), we obtain a(uh, uh) +
-'n
®(qhuh) dx < a(uh, rhv) +
•'a
V c e 2(Q). (2.95)
F r o m (2.85), (2.89), a n d L e m m a 2.1 we have a(u*, u*) + ("
Vȣ
Then in the limit in (2.95) we obtain a(u*,
u*) + j{u*) < a(u*, v) + j(v) - L(v - u*\
V n e 3>(Q).
(2.96)
From Fatou's lemma applied to (2.84) and (2.91) we obtain
(2.97)
Then from (2.96) and (2.97) it follows that u* satisfies a(u*, v-u*)+
j(v) - j(u*) > L(v - u*),
V » e @(Q), u* e V, d>(u*) 6 L'(fl). (2.98)
We now take v e V n L co (n); from Lemma 2.5 it follows that there exists a sequence (»„)„, vn e @(ti), such that lim vn = v strongly in V, (2.99) lim vn = v in Lm(Q) weak *.
(2.100)
126
IV Applications of Elliptic Variational Inequality Methods
From (2.98) we have a(u*, vn - «*) + j{vn) - j(u*) > L{vn - u*),
V n, u* e V, O(u*) 6 L\Q).
(2.101)
From (2.99) we obviously have lim a(u*, vn — u*) = a(u*, v — u*), n-*oo
lim L(vn — u*) = L(v — u). Since un -* v in the weak* topology of L°°(£X), we have a constant C such that K l U n , < C,
Vn.
(2.102)
Moreover, for some subsequence, (2.99) implies lim vn = v a.e. in O.
(2.103)
From (2.103) we obtain <$(vn) ->
n—w •'n
Therefore, taking the limit in (2.101), we obtain a(u*, v-u*)+
j(v) - j(u*) > L(v - «)*,
'iveVn
L°°(n), u* e V, <£(«*) e L\O).
(2.104)
Since from Lemma 2.2 we know that (2.104) is equivalent to (n), we have proved that u* = u, where u is the solution of (71). From the uniqueness of the solution of (n), it follows that the whole sequence (uh)h converges to u. (3) Strong convergence of(uh)h. From the K-ellipticity of a(-, •) and from the variational inequality satisfied by uh, we obtain all"* - "IIK + ;»(«*) ^ «(«* - «. «* - «) + jh(uh) < -a(uh, u) + a(u, u) + a(uh, uh) -a(u, uh) +jh(uh) < - a(uh, u) + a(u, u) + a(uh, rhv) + jh(rhv) - L(rhv - uh) -a(u,uh), V»e®(Q). (2.105) Using the various convergence results of part (2), from (2.105) we obtain X") ^ lim inf jh(uh) < lim inf(a||uft - u\\l + jh(uh)) < l i m s u p ( a | | « f c - u\\y + jh(uh)) < a(u, v - u) + }(v) - L(v - u),
V c e ®(Q). (2.106)
As in part (2), using the density of &{il) in L°°(Q) n V (for the strong topology of V and the weak * topology of L^fi)), we find that (2.106) also holds for all v e VnLx(£T). Then j(u) < lim inf jh(uh) < lim inf(a||wft - u\\v + jh(uh)) < lim sup(a||Mft - u\\r + jh(uh)) < a(u, xnv - u) + j(znv) - L{xnv - v),
VPEF. (2.107)
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
127
Using the properties of xnv, it then follows from (2.107), by taking the limit as n -» co, that (2.106) also holds for all v e V. Hence, by taking v = u, we obtain j(u) < lim inf jh(uh) < lim i n f ( a K - u\\v + jh(uh)) < lim sup(a||u h - u\\v + jh(uh))
<j(u\
which implies lira jh(uh)
=j(u),
lim \\uh - u\\v = 0.
This proves the theorem, modulo the proof of (2.88).
D
We now prove (2.88). Lemma2.6. We have Vp, 1 < p < oo, \\qhvh - vh\\Lm) < f/i||V^ ILp^xt^n) where qh, vh are as before. PROOF. We use the notation of Sec. 2.5.1. We have (see Fig. 2.5) | vh(M) - qh vh(M) | = | vh(Md ~ vh(M) |,
V M e ilt n T.
(2.108)
But since vh\T e Pv we have vh(M) = v^Mt) + M~M -\7vh,
y
Meilt^T,
from which, combined with (2.108), it follows that \qhvh(M) - vh(M)\ < | But from the definition offt,we have \MtM\ < fft, V T, so that we finally have \qhvh(x) - vh(x)\ <±h\Vvh(x)\ a.e. in Q,V vheVh. This implies hhVh - vh\\ma)
This proves the lemma.
<
ih\\Vvh\\mil)xLP(il).
•
Remark 2.4. We have not considered the problem of error estimates. This problem is discussed in Glowinski [5].
Figure 2.5
M,
128
IV Applications of Elliptic Variational Inequality Methods
Remark 2.5. The numerical analysis of problems like (P), but with much stronger hypotheses on a( •, •), (f>, and / , is considered in Ciarlet, Schultz, and Varga [1].
2.6. Iterative methods for solving the discrete problem 2.6.1. Introduction In this section we briefly describe some iterative methods which may be useful for computing the solution of (Ph) (and (nh)). Actually most of these methods can be extended to other nonlinear problems. Many of the methods to be described here can be found in Ortega and Rheinboldt [1]. A method based on duality techniques will be described in Chapter VI. 2.6.2. Formulation of the discrete problem Here we are using the notation of the continuous problem. Taking as unknowns the values of uh at the interior nodes of 3~h, the problem (P,,) reduces to the finite-dimensional nonlinear problem Au + D0(u) = f,
(2.109)
where A is a N x N positive definite matrix, D is a diagonal matrix with positive diagonal elements dt and where u = {uu ..., uN} e UN, f e UN, 0(u) e UN with (>(u)); =
u"
+1
= u" --
N
eU
pS~\Au" +
Clj) is given by
given,
D(Xu") --f),
(2.110) P > 0.
(2.111)
In (2.111), S is a symmetric positive-definite matrix; a canonical choice is S = identity. But in most problems it will yield a slow speed of convergence. If A is symmetrical, the natural choice is S = A, and if A # A*, we can take S = (A + A*)/2. For the convergence of u" to u (where u is the solution of (2.109)), it is sufficient to have
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
129
Remark 2.7. In each specific case, p has to be determined; this can be done theoretically, experimentally, or by using an automatic adjustment procedure which will not be described here. Remark 2.8. Let us define g" by g" = Au" + D
(2.112)
Suppose A = A*; then if we use (2.110), (2.112) with pn defined by J(u--pI1S-1r)^J(u"-pS-1g")>
VpeR,
pneU,
(2.113)
the resulting algorithm is, for obvious reasons, called the steepest descent method. This algorithm is convergent for > £ C°(K). We observe that at each iteration the determination of pn requires the solution of a one-dimensional problem; for the solution of such one-dimensional problems, see Householder [1], Polak [1], and Brent [1]. Remark 2.9. At each iteration of (2.110), (2.111) or (2.110), (2.112), we have to solve a linear system related to S. Since S is symmetric and positive definite, this system can be solved using the Cholesky method, provided the S = LL* factorization has been done. From a practical point of view, it is obvious that the factorization of S will be made in the beginning once and for all. Then at each iteration we only have to solve two triangular systems, which is a trivial operation. 2.6.4. Newton's method The Newton's algorithm is given by (for sufficient conditions of convergence, see Ortega and Rheinboldt [1]) u° £ UN given, u"
+1
= (A + D f t O r ^ D ^ u " ) ! ! " - D0(u") + f),
(2.114) (2.115)
where >'(v) denotes the diagonal matrix
Since
130
IV Applications of Elliptic Variational Inequality Methods
which avoid this drawback have been designed. In this regard, let us mention, among others, Broyden [1], [2], Dennis and More [1], Crisfield [1], Matthies and Strang [1], O'Leary [1], and also the references therein; most of the above references deal with the so-called quasi-Newtori's methods. Remark 2.11. The choice of u° is very important when using Newton's method. 2.6.5. Relaxation and over-relaxation methods. In this section we shall discuss the application of relaxation methods for solving the specific problem (2.109); the algorithms to be described below are, in fact, particular cases of a large family of algorithms to be discussed in Chapter V. In this section we use the following notation:
f = {/i,/2, • • • , / * } •
Since A is positive definite, we have au > 0, V i = 1, 2 , . . . , N. Here we will describe three algorithms. ALGORITHM 1
u°eUN given;
(2.116)
n+1
then with u" known, we compute u +i
auu?
1
, component by component, using
= ft - E«u"" + 1 - l>u«".
+ d^mr )
u 1 + 1 = u1 + co(u1+1
-«?),
i=l,2,...,N.
(2-117) (2.118)
Since an > 0, > e C°(U) and
ALGORITHM
anu1+1 + Jf, \
- I fly«J+' - £ ayMj) for i = 1, 2,..., JV. ]
}>i
I
(2.119)
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
131
Remark 2.12. If co = 1 and/or cj> is linear, the two algorithms coincide. In the general case the convergence of (2.116), (2.119) seems to be an open question. However, from numerical experiments it seems that the second algorithm is more "robust" that the first, perhaps because it is more implicit. Furthermore, it can be used even if cf> is defined only on a bounded or semibounded interval ]a, /?[ of U such that >(<*) = - oo, >(/?) = + oo; in such a case, if c/> e C°(]a, /?[) and (j) is increasing, (2.109) has still a unique solution, but the use of (2.116)— (2.118) with co > 1 may be dangerous. Remark 2.13. If c/> e C\U), an efficient method for computing u"+ i in (2.117) and M"+ 1 in (2.119) is the one-dimensional Newton's method. Let g £ C^IR). In this case, Newton's algorithm for solving the equation g(x) = 0 is x° £ U given,
(2.120)
x "+i = x "_^gl.
(2.121)
If in the computation of u"+1 and M"+ 1, we use only one iteration of Newton's method, starting from u", then the resulting algorithms are identical and we obtain: u°eUN given,
(2.122)
au + fl; CO (M; )
(2.123) In S. Schechter [1], [2], [3], sufficient conditions for the convergence of (2.122), (2.123) are given. Remark 2.14. If co > 1 (resp., co = 1, co < 1), the previous algorithms are overrelaxation (resp., relaxation, under relaxation) algorithms. Remark 2.15. In Glowinski and Marrocco [1], [2], we can find applications of relaxation methods for solving the nonlinear elliptic equations modeling the magnetic state of electrical machines. 2.6.6. Alternating-direction methods. In this section we take p > 0. Here we will give two numerical methods for solving (2.109). First method u°eUN given;
(2.124)
132
IV Applications of Elliptic Variational Inequality Methods
knowing u", we compute u" + 1 / 2 by ) + f; then we calculate u" pu"
+1
+1
(2.125)
by + D)(u"+1) = pu" + 1/2 - Au" +1/2 + f.
(2.126)
For the convergence of (2.124)-(2.126), see R. B. Kellog [1]. Second method u°eUN given;
(2.127)
knowing u", we compute u" + 1 / 2 fey pu »+i/2 +
Au n + 1 / 2 = pu" - D0(u") + f;
(2.128)
= pu" - Au"+ 1/2 + f.
(2.129)
then we calculate u" +1 by
Using the results of Lieutaud [1], we can prove that, for all p > 0, u" + 1 / 2 and u" converge to u if we suppose that A and > satisfy the hypotheses given in Sec. 2.6.2; the same convergence result holds if we exchange the role of A and in (2.128), (2.129). Remark 2.16. At each iteration we have to solve a linear system whose matrix is constant, since we use a constant step p. This is an advantage from the computational point of view (cf. Remark 2.9). We also have to solve a nonlinear system of N equations, but in fact these equations are independent of each other and reduce to N nonlinear equations in one variable, which can be solved easily. 2.6.7. Conjugate gradient methods. In this section we assume A = A* (if A ^ A*, we can also use methods of conjugate gradient type). For a detailed study of these methods, we refer to Polak [1], Daniel [1], and Concus and Golub [1]. If the functional J defined in Remark 2.6 is not quadratic (i.e., cf> is nonlinear), several conjugate gradient methods can be used. Let us describe two of them, whose convergence is studied in Polak [1]. Let J be given by J(v) = i(Av, v) + X dtdXvd - (f, v), i=i
where CE>(t) = f/0 C/>(T) di, (j) being a nondecreasing continuous function on U with >(0) = 0. Let S be a N x N symmetric positive-definite matrix.
2. Theoretical and Numerical Analysis of Some Mildly Nonlinear Elliptic Equations
133
First method (Fletcher-Reeves) u°eUN given; g° = S" W
(2.130)
+ D0(n°) - f), w° = g°.
(2.131) (2.132)
Then assuming that a" and w" are known, we compute un+1 by u" + 1 = u " - p n w " ,
(2.133)
where pn is the solution of the one-dimensional minimization problem J(u" - pnwn) < J(u" - pw"),
V p e U, pneU.
(2.134)
Then g"+1 = S'HAu""1"1 + D0(u"+1) - f),
(2.135)
and compute w" +1 by w" +1 = g" +1 +2 n w",
(2.136)
where (Sen+1 K
s" +1 ) (2137)
(sg",g«)
Second method (Polak-Ribiere). This method is like the previous method, except that (2.137) is replaced by g"+1g"+1-g") )
•
(2138)
Remark 2.17. For the computation of pn in (2.134), see Remark 2.8 (and also Shanno [1]). Remark 2.18. From Polak [1] it follows that if (f> is sufficiently smooth, then the convergence of the above algorithms is superlinear, i.e., faster than the convergence of any geometric sequence. Remark 2.19. The above algorithms are very sensitive to roundoff errors; hence double precision may be required for some problems. Moreover, it may be convenient to periodically take w" = g" (see Powell [2] for this restarting procedure problem). Remark 2.20. At each iteration we have to solve a linear system related to S; Remark 2.9 still applies to this problem.
134
IV Applications of Elliptic Variational Inequality Methods
2.6.8. Comments. The methods of this section may be applied to more general nonlinear systems than (2.109). They can be applied, of course, to finite-dimensional systems obtained by discretization of elliptic problems like — V • (ao(x)Vj<) + /? • VM + (j)(x, u) = f in Q,
plus suitable boundary conditions
where, for fixed x, the function t -> 4>(x, t) is continuous and nondecreasing on U.
3. A Subsonic Flow Problem 3.1. Formulation of the continuous problem Let Q be a domain of UN (in applications, we have N = 1,2,3) with a sufficiently smooth boundary T. Then the flow of a perfect compressible irrotational fluid (i.e., V x v = 0, where v is the velocity vector of the flow) is described by OinQ,
[(y + D/(y -
(3.1)
J
with suitable boundary conditions. Here: • • • • •
(j> is a potential and V0 is the velocity of the flow; p{4>) is the density of the flow; p0 is the density at V> = 0; in the sequel we take p0 = 1; y is the ratio of specific heats (y = 1.4 in air); C^ is the critical velocity.
The flow under consideration is subsonic if < C # everywhere in Q.
(3.3)
If | V(j> | > C^ in some part of Q, then the flow is transonic or supersonic, and this leads to much more complicated problems (see Chapter VII for an introduction to the study of such flows). Remark 3.1. In the case of a subsonic flow past a convex symmetrical airfoil and assuming (see Fig. 3.1) that v^ is parallel to the x-axis (Q is the complement of the airfoil in IR2 and d(j)/dn\r = 0), H. Brezis and Stampacchia [1] have proved that the subsonic problem is equivalent to an EVI of the first kind in the hodograph plane (see Bers [1] and Landau and Lifschitz [1] for the hodograph transform). This EVI is related to a linear operator, and the corresponding convex set is the cone of non-negative functions.
3. A Subsonic Flow Problem
135
Figure 3.1
In the remainder of Sec. 3 (and also in Chapter VII), we shall work only in the physical plane, since it seems more convenient for the computation of nonsymmetric and/or transonic flows. For the reader who is interested by the mathematical aspects of the flow mentioned above, see Bers [1] and Brezis and Stampacchia [1]. For the physical and mechanical aspects, see Landau and Lifschitz [1]. Additional references are given in Chapter VII.
3.2. Variational formulation of the subsonic problem
Preliminary Remark. In the case of a nonsymmetric flow past an airfoil (see Fig. 3.2) the velocity potential has to be discontinuous, and a circulation condition is required to ensure the uniqueness (modulo a constant) of the solution of (3.1). If the airfoil has corners (as in Fig. 3.1), then the circulation condition is related to the so-called Kutta-Joukowsky condition from which it follows that for a physical flow, the velocity field is continuous at the corners (like 0 in Fig. 3.2). For more information about the Kutta-Joukowsky condition, see Landau and Lifschitz [1] (see also Chapter VII). For the sake of simplicity, we shall assume in the sequel that either Q is simply connected, as is the case for the nozzle of Fig. 3.3, or, if Q is multiply connected, we shall assume (as in Fig. 3.1) that the flow is physically and geometrically symmetric, since in this case the Kutta-Joukowsky condition is automatically satisfied. In the sequel we assume that the boundary conditions associated with (3.1), (3.2) are the following: g0overTr0,p-^
l
= 9,
Figure 3.2
(3-4)
136
IV Applications of Elliptic Variational Inequality Methods Figure 3.3
n
where F o , F t c F with F o n F x = 0 , F o u F t = F. Then the variational formulation for the flow problem (3.1), (3.2), (3.3), (3.4) is • Vv dx
gtvdr,
VveV0,
(3.5)
where (3.6)
Vto={veH\n),v\ro
= g0}.
(3.7)
If g0, g1 are small enough, we can prove that (3.5) has a solution such that <M
C # a.e..
When solving a practical flow problem, we may not know a priori whether the flow will be purely subsonic or not. Therefore, instead of solving (3.5), it may be convenient to consider (and solve) the following problem:
~4>)dx > f dl{v - 4>)dT,
VveKd,
4>eKd, (3.8)
where (3.9) The variational problem (3.8), (3.9) is an EVI of the first kind, but we have to observe that unlike the EVI's of Chapters I and II, it involves a nonlinear partial differential operator, namely, A defined by A{4>) = - V Remark 3.2. In practical applications we shall take 6 as close as possible to C*. Remark 3.3. Problem (3.8), (3.9) appears as a variant of the elasto-plastic torsion problem of Chapter II, Sec. 3.
3. A Subsonic Flow Problem
137
3.3. Existence and uniqueness properties for problem (3.8) In this section we shall assume that meas(T 0 ) > 0. To prove that (3.8) is well posed, we will use the following: Lemma 3.1. The function
is convex if ^ E [0, C^\, concave if
and strictly convex if £ £ [0, C^[_. EXERCISE 3.1. Prove Lemma 3.1.
We can now prove: Theorem 3.1. Assume that Q is bounded and that g0, gx are sufficiently smooth and small. Then (3.8) has a unique solution. PROOF. Since Q is bounded, and if g0 is sufficiently smooth and small, we observe that K6 is a closed, convex, and nonempty bounded subset of H1(Q.) (consisting of uniformly Lipschitz continuous functions). Define J ( ) by
From Lemma 3.1 it follows that J(-) is strictly convex over Ks. It is easy to check that J(-) is continuous and Gateaux-differentiable over Kd with (J'(v), w) = f p(v)Vv -Vwdx-
\ QiWdT.
(3.11)
Since Kd is a closed convex nonempty subset of Hl(Q) and J ( ) is continuous and strictly convex over Kt, from standard optimization theory in Hilbert spaces (see, e.g., Cea [1], [2]), it follows that the minimization problem J(u)<J(v),
VceXj,
ueKd,
(3.12)
has a unique solution. Moreover, since J(-) is differentiable, the unique solution of (3.12) is characterized (see Cea [1], [2]) by (J'(«), v - u) > 0,
VceKj,
from (3.11), this completes the proof of the theorem.
UEKS;
138
IV Applications of Elliptic Variational Inequality Methods
Remark 3.4. Suppose that F o = 0. Then defining K5 by Ks={ve
H'iQ), \Vv\<3
a.e., v(x0) = v0}
with x 0 e Q and v0 arbitrarily given, we can prove that if Q is bounded and gfx is sufficiently smooth, then
f p(
VveKd,
Jn Jr has a unique solution (if (f> is a solution of (3.13); then cf> + C is the unique solution of the similar problem obtained by replacing v0 by v0 + C). EXERCISE
3.2. Prove the statement of Remark 3.4.
Remark 3.5. In all the above arguments we assumed that Q. is bounded. We refer to Ciavaldini, Pogu, and Tournemine [1] which contains a careful study of the approximation of subsonic flow problems on an unbounded domain Q^ by problems on a family (£!„)„ of bounded domains converging to Qx (actually they have obtained estimates for (p^ — >„). The above EVFs will have a practical interest if we can prove that in the cases where a purely subsonic solution exists, then for 3 large enough it is the solution of (3.8); actually this property is true and follows from: Theorem 3.2. Assuming the same hypotheses on Q, g0, qx as in Theorem 3.1, and that (3.1), (3.2), (3.4) has a unique solution in H\Q.) with |V0|<<5 O
(3.14)
then (f> is a solution o/(3.8), (3.9), V 3 e [<50, C^Q Conversely, if the solution of (3.8), (3.9) is such that \ V $ | < 60 < 3 < C* a.e., then (f> is a solution of (3.1), (3.2), (3.4). PROOF. (1) Let <j> e H^Q) satisfy (3.1), (3.2), (3.4), and (3.14). If v e Vo, then using Green's formula, it follows from (3.1), (3.2), (3.4) that f p(0)V0 • Vv dx = f QivdT,
MveVo.
(3.15)
From (3.4), (3.15) and from the definition of Vgo it follows that f p{4>)V4> • V(v - 4 > ) d x = \ 9 l ( v - e Kd a Vgo, V 5 e [c50, C # [, it follows from (3.16) that
V v e VSo.
J
V(B -
9i(v
- <j>) dT,
VveKs,
<j> e Kd
*'ri
if 5 e [50, C^C; therefore 0 is the solution of the EVI (3.8), (3.9), V 5 e [<50, CJ_.
(3.16)
3. A Subsonic Flow Problem
139
(2) Define U <= Vo by U = {ve C°°(fi), v = 0 in a neighborhood of F o }. Then, if we suppose that F is sufficiently smooth, we find that the closure t/ fll(n) of U in H\il) obeys t7 HI(fi) = Vo.
(3.17)
We assume that for 5 < C^, (3.8) has a solution such that \V
(3.18)
Then V t> e [/, and for f > 0 sufficiently small, <$> + tv e Kd. Then replacing t> by <> / + ft) in (3.8) and dividing by t, we obtain f p(<^>)V0 • Vv dx > \ c,
r,
•lr1
JQ
V»E[/,
which implies f
f p(4>)V4> • Vv dx =
gtv dT,
V v e U.
(3.19)
Since &(Q.) a U, it follows from (3.19) that f p(<^))V0 • Vt; Jx = 0,
V v e @(Q),
(3.20)
i.e., which proves (3.1). Assuming (3.1) and using Green's formula, we obtain f pW>)V> • VP dx = f p — v dT,
\/veU.
(3.21)
Using (3.17) and comparing with (3.19), we obtain
P
tn
i.e., (3.4), which completes the proof of the theorem.
Remark 3.6. A similar theorem can be proved for the problem mentioned in Remark 3.4. 3.4. Comments
The solution of subsonic flow problems via EVFs like (3.8) or (3.13) is considered in Ciavaldini, Pogu, and Tournemine [2] (using a stream function approach) and in Fortin, Glowinski, and Marrocco [1], Iterative methods for solving these EVFs may be found in the above references and also in Chapter VI of this book.
CHAPTER V
Relaxation Methods and Applications1
1. Generalities The key idea of relaxation methods is to reduce, using some iterative process, the solution of some problems posed in a product space V = YW= 1 ^ (minimization of functional, solution of systems of equations and/or inequalities, etc.) to the solution of a sequence of subproblems of the same kind, but simpler, since they are posed in the Vt. A typical example of such methods is given by the classical point or block Gauss-Seidel methods and their variants (S.O.R., S.S.O.R., etc.). For the solution of finite-dimensional linear systems by methods of this type, we refer to Varga [1], Forsythe and Wasow [1], D. Young [1] and the bibliographies therein. For the solution of systems of nonlinear equations, we refer to Ortega and Rheinboldt [1], Miellou [1], [2], and the bibliographies therein. For the minimization of convex functionals by methods of this kind, let us mention S. Schecter [1], [2], [3], Cea [1], [2], A. Auslender [1], Cryer [1], [2], Cea and Glowinski [1], Glowinski [6], and the bibliographies therein. The above list is far from complete. The basic knowledge of convex analysis required for a good understanding of this chapter may be found in Cea [1], Rockafellar [1], and Ekeland and Temam [1].
2. Some Basic Results of Convex Analysis In this section we shall give, without proof, some classical results on the existence, uniqueness, and characterization of the solution of convex minimization problems. Let (i) V be a real reflexive Banach space, V* its dual space, (ii) K be a nonempty closed convex subset of V, 1
In this chapter we follow Cea and Glowinski [1] and Glowinski [6].
2. Some Basic Results of Convex Analysis
141
(iii) J o : V -> U, be a convex functional Frechet or Gateaux differentiable2 on V, (iv) Jt: V -+M, be 3 a proper l.s.c. convex functional. We assume that K n D o m ^ ) # 0 , where Don^Jj) = {v\v e V, J^v) e K}. We define J:V^U
by J = J0 + Ji and assume that lim
J(v)=
+oo.
(2.1)
|M|-> + 00 VEK
Under the above assumptions on K and J, we have the following fundamental theorem. Theorem 2.1. The minimization problem J(u) < J(v),
VceK,
ueK,
(2.2)
has a solution characterized by <J'o(u), D - u> + JiOO - Jtiu) > 0,
VceJC,
M e X.
(2.3)
4
77HS solution is unique if J is strictly convex.
Remark 2.1. If K is bounded, then (2.1) may be omitted. Remark 2.2. Problem (2.3) is a variational inequality (see Chapter I of this book). Let us now recall some definitions about monotone operators. Definition 2.1. Let A: V -> V*. The operator A is said to be monotone if
0,
Vu,veV,
and strictly monotone if it is monotone and
(A(v) - A(u), v - u) > 0,
Vu,veV,
u # v.
Let F: V -> R; the Gateaux-differentiability property means that
where <•, •> denotes the duality between V* and V and F ^ e f 7 * ; F'(v) is said to be the Gateaux-derivative (or G-derivative) of F at v. Actually, we shall very often use the term gradient when referring to F'. !
R = Ru { + 00} u {-oo}.
* i.e., J(tv + (1 - t)w) < U(v) + (1 - t)J(w), V t e ]0,1[, V v, w e Dom(J), v * w.
142
V Relaxation Methods and Applications
We shall introduce the following proposition which will be very useful in the sequel of this chapter. Proposition 2.1. Let F: V —> R be G-differentiable. Then there is equivalence between the convexity of F (resp., the strict convexity of F) and the monotonicity (resp., the strict monotonicity) of F'. To prove Theorem 2.1 and Proposition 2.1, we should use the following: Proposition 2.2. If F is G-differentiable, then F is convex if and only if F(w) - F(v) >
Vv,weV.
(2.4)
3. Relaxation Methods for Convex Functionals: Finite-Dimensional Case 3.1. Statement of the minimization problem. Notations With respect to Sec. 2, we assume that V = UN, with v = {»!,..., vN},
Vi
eU,
1 < i < N.
The following notation will be used in the sequel: JV
(M, V) = £ UiVh
\\V\\ = J(V, V).
i= l
We also assume that K = {v\v e UN, Vi e Ki = lat, bj, at < bh 1 < i < N},
(3.1)
where the a{ (resp., theft,-)may take the value — oo (resp., + oo); K is obviously a nonempty closed convex subset of RN. Furthermore, we assume that J(v) = J0(v)+ YhiVil
(3-2)
where Jo e C\UN) and is strictly convex and where the j , e C°(U) and are convex, 1 < i < N. We note that we are not assuming the differentiability of the ji. Finally, we assume that lim
J(v) = +oo.
(3.3)
|| V || -* + 00
Then, under the above assumptions and from Theorem 2.1, it follows that the problem J(u) < J(v) VveK, ueK, (3.4)
3. Relaxation Methods for Convex Functionals: Finite-Dimensional Case
143
has a unique solution characterized by (J'o(u), v-u)+
£ (/i(»i) -;<(«>)) 2: 0,
VceK,
u e X,
(3.5)
t=i
where J'0(v) denotes the gradient of Jo at v.
3.2. Description of the relaxation algorithm To solve (3.4) we can use the following relaxation algorithm: «° € K arbitrarily given;
(3.6)
n+1
then, u" being known, we compute u
, component by component, by
<J{u\+\...,utl,vhu1^,..),
VvteKh
ur'eK,,
(3.7)
where 1 < i < N. From the above assumptions on K and J, and from Theorem 2.1, we obtain: Proposition 3.1. Each subproblem (3.7) has a unique solution which is characterized by ^ « \ . . . , u r \ « ?
+
j , • • .X»( - «?+ x) + ;,<»i) -7i(«7 + ') > o,
u1+1eKi.
VveK(,
(3.8)
Remark 3.1. To compute u"+1, we can proceed as follows. First we compute u" + 1 by solving J^H!
,...,«,-_!,«,-
, Ui+1, . . . ) S Jyu-L
, . . . , M ; - i , Vt, Ui+1, . . . ) ,
\fvteU,
M?+16lR.
(3.9)
Then we project u"+1 on Kt to obtain w"+ i; so «7+
L
= PKi(u1+ J ) = Max(a ; , Minfo, S?+ x)).
(3.10)
If 7; is differentiable, then w"+x is the solution of the single-variable equation ^ ( M - 1 + 1 , . . . , « ? _ + 1 1 , S ? + 1 ) H 7 + ! , . . . ) + ^ ( M ? + 1 ) = 0.
(3.11)
Equation (3.11) can be solved by various methods (see, for example, Householder [1]).
144
V Relaxation Methods and Applications
3.3. A lemma on the monotonicity of the gradient of strictly convex functionals To prove the convergence of algorithm (3.6), (3.7), we shall use the following: Lemma 3.1. Assume that F e C^IR^) and is strictly convex. Then F is uniformly convex on the bounded sets of UN, i.e., V M > 0, 3 5M: [0, 2M] -> U+, continuous, strictly increasing, and such that dM(0) = 0, (F'(v) - F'(u), v-u)> N
Vu, veR ,
(3.12) 3M(\\v -
\\u\\ < M,
M||)
|M|<M,
(3.13)
and F(v) > F(u) + (F'(u), v-u) \fu,veUN,
\\u\\ < M,
+ &M(\\v
- M||),
\\v\\ < M.
(3.14)
PROOF. Let BM = {v\ve UN, \\v\\ < M}. For x e [0, 2M], we define d^, by ^(t)=
inf
(F(v) ~ F'(u), v - u).
||v—u|| = x u,veB\f
(?•*••>)
From the definition of 5% it follows that S°M(0) = 0
(3.16)
and (F'(v)
- F'(u),
v - u ) > 5°M(\\v
- u\\),
Vu,ve BM.
(3.17)
£
Let T 2 ] 0 , 2M]. From the continuity of {u, v} -* (F'(v) — F'(u), v — u) and from the compactness of BM x BM, it follows that there exists at least one pair {u2, v2} realizing the minimum in (3.15). Then <52f(T2) = (F'(v2) - F'(u2), v2 - u2\ and <5°,(T2) > 0
from the strict monotonicity o f f (cf. Sec. 2, Proposition 2.1). Let x^ 6 ]0, T 2 [. We define w e ] u 2 , f 2 [ by w = u2 -\
(v2 — u2).
Since 0 < x1/x2 < 1, from the strict monotonicity o f f it follows that (F'(v2) - F'(u2), v2 - u2) > [F'(u2 + -(V2\
\
T
2
U2)) - F'(u2), v2 - u2). /
/
3. Relaxation Methods for Convex Functional: Finite-Dimensional Case
145
This implies (F'(v2) - F'(u2), v2 - u2) > - (F'(w) - F\u2), w - u2) > (F'(w) - F'(u2), w - u2). (3.18) Since j|w — u2\\ = z1, (3.18) in turn implies
Applying (3.17) to {u + t(v - u), u}, it follows that (F'(u + t(v - u)), v-u)>
(F'(u), « - « ) + | ^Ktlli; - u\\),
V t e ]0, 1]. (3.19)
From the continuity of F , it easily follows that lim - 5°M(?) = 0.
(3.20)
Then from (3.20) it follows that (3.19) could be extended at t = 0. Integrating (3.19) on [0, 1], it follows that F(v) - F(u) > ( F ( » ) , v-u)+
f 5°M(t\\v - ii||) - . Jo
(3.21)
'
We also have F(u) - F(v) > (F'(v), u-v)+
f d°M(t\\v - u||) - .
(3.22)
Then, by summation of (3.21), (3.22), we obtain (F'(v) - F'(u), v - u) > 2 f d°Jt\\v - M||) Jo t (•\\v-u\l
= 2 Jo
j
SUs)~s
0.23)
Therefore the function 5M defined by 5M{x) = 2 CdUs) Jo s
(3-24)
has the required properties. Furthermore, (3.14) follows from (3.21) and from the definition of 5M. • Remark 3.2. The term forcing function is frequently used for functions such as dM (see Ortega and Rheinboldt [1]).
146
V Relaxation Methods and Applications
3.4. Convergence of algorithm (3.6), (3.7) We have: Theorem 3.1. Under the above assumptions on K and J, the sequence (u")n defined (3.6), (3.7) converges, V u° e K, to the solution u of (3.4). PROOF. For the sake of simplicity, we have split the proof into several steps. Step 1. We shall prove that the sequence J(u") is decreasing. We have J(u") - J(u"+1) = £ ( J « 1 , . . . , uill
ul...)-
J « \ ..., uT \ u?+!,.. •)). (3.25)
Since u? e Kt, it follows from (2.4), (3.8) that, V i = 1 , . . . , N, we have J(ul+ \ . . . , utl
J(u\+ \...,
ul,...)-
uT \ u% „ ...)
+
J0(u\ \ ..., « T \ u1+ u ...) + jiui) - JM+1)
= J 0 ( « r ' , . . . , utl, ul...)-
(3.26)
Then (3.26) combined with (3.25) implies J(u")> J(un+l),
Vn>0.
(3.27)
Moreover, since J satisfies (3.3), it follows from (3.27) that there exists a constant M such that IMI <M, \\u"\\ <M, ||{M-1+1.---,«I
+ 1
V«,
,«?+I,...}II^M,
V(=l,...,Af,
VH.
(3.28)
Step 2. F r o m (3.8), (3.14), (3.25), (3.26), and (3.28), it follows that
J(u") - J(u"+!) > \ X 5M{ | a? + 1 - H? |).
(3.29)
The sequence J(u") is decreasing and bounded below by J(u), where w is the solution of (3.4). Therefore the sequence J(u") is convergent, and this implies lim (J(un) - J(un+1)) = 0.
(3.30)
From (3.29), (3.30), and the properties of SM, it follows that lim ( « " - u" + 1 ) = 0.
(3.31)
Step 3. Let u be the solution of (3.4). Then it follows from (3.15), (3.28) that C W + 1 ) - J' 0 (u), u"+i -u)>
5MQ\u"+1
- u\\),
which implies (J'0(u"+1) - J'0(u\ u"+l -u) + J^u"*1) - J^u) > JW+1) - JM
+ <5M(||u"+1 - M||). (3.32)
3. Relaxation Methods for Convex Functionals: Finite-Dimensional Case
147
Since u is the solution of (3.4) and un+' e K, we have (cf. (3.5)) (J'0(u), u" + l - a) + Jt{u"+i) - JM
> 0,
which, combined with (3.32), implies (io(«" + 1 ),«" + 1 - u) + JW+1)
- JM
> dM(\\u"+1 - till).
(3.33)
Relation (3.33) implies (J' 0 (u" +1 ), un+1 - u) + - W
N
) - ^(u)
tdl +1
+
1
-«II),
1
(3-34)
n
where fl? ' = { u f , . . . , uT \ u V i = 1 , . . . , N,
i+ u
...}. Since ut e Kh
+
^
it follows from (3.8) that,
' ) - yiC***) < o.
(3.35)
Therefore (3.34) and (3.35) show that
f (""+1}" af ("r 1 } r r * ~~Ui) - dMu"+1 ~u|l)- (3-36) «" +1 - u1+1\\ < \\un+1 - uunn\\\\, it follows from (3.31) that V i = 1 , . . . , N, we have Since ||«" lim ( w " + 1 - « r 1 ) = 0.
(3.37)
Since J'o 6 C°(RN), J'o is uniformly continuous on the bounded subsets of U". This property, combined with (3.37), implies, V i = 1 , . . . , N, = 0.
lim
(3.38)
n~* + oo
Therefore, from (3.28), (3.36), (3.38), and the properties of8M, it follows that lim ||u" - u\\ = 0, n-* oo
which completes the proof of the theorem.
•
3.5. Various remarks Remark 3.3. We assume that K = UN and that J = Jo (i.e., Jx == 0), where jo(v)
= ±(Av, v) - (b, v),
where beUN and
A is an N x N symmetrical positive-definite matrix.
148
V Relaxation Methods and Applications
The problem (3.4) associated with this choice of J and K obviously has an unique solution characterized (cf. (3.5)) by Au = b.
(3.39)
If we apply the algorithm (3.6), (3.7) to this particular case, we obtain u° e UN, arbitrarily given;
- L.«y«r - I «u«"j.
(3.40)
1 ^ i < N.
(3-41)
The algorithm (3.40), (3.41) is known as the Ganss-Seidel method for solving (3.39) (see, e.g., Varga [1] and D. Young [1]). Therefore, when A is symmetric and positive definite, optimization theory yields another proof of the convergence of the Gauss-Seidel method through Theorem 3.1. Remark 3.4. From the above remark it follows that the introduction of overor under-relaxation parameters could be effective for increasing the speed of convergence. This possibility will be discussed in the sequel of this chapter. L e t F : F - + U. We define D(F)= {v\veV,\F(v)\ < +oo}.
(3.42)
If F is convex and proper, then D(F) is a nonempty convex subset of V. Remark 3.5. If in Sec. 3.1 we replace the conditions j , € C°(U) and jt convex, V i = 1 , . . . , N, by ji: U -> U is convex, proper, and l.s.c. and we assume Kt n DO',) ^ 0, V i = 1 , . . . , N, then the other assumptions being the same, (3.4) is still a well-posed problem and (3.5) still holds. Moreover, the algorithm (3.6), (3.7) could be used to solve (3.4), and the convergence result given by Theorem 3.1 would still hold. Remark 3.6. We can complete Remark 3.5 in the following way. We take jt as in Remark 3.5 and assume Jo strictly convex, proper, and l.s.c., D(J0) is an open set of UN and J o e C1(/>(J0))- T h e n , if D(J) n K ± 0 and if lini||i;||_ + 00 J(v) = +oo, problem (3.4) is well posed and (3.5) still holds. Moreover, algorithm (3.6), (3.7) could be used to solve (3.4). Remark 3.7. A typical situation in which algorithm (3.6), (3.7) could be used isK = UN, Jo as in Remark 3.3 and J^v) = Vf=1 a ; |i;,|,a; > 0 , V i = l,...,N.
3. Relaxation Methods for Convex Functionals: Finite-Dimensional Case
149
3.6. Some dangerous generalizations In this section we would like to discuss some of the limitations of the relaxation methods. 3.6.1. Relaxation and nondifferentiable functionals We consider K = U2 and J e C°(U2), strictly convex, defined by
J(v) = Wi + vl) + K - v2\ - (»! + v2); J is nondifferentiable on the line v1 = v2. The unique solution of J(u) = MinJ(v)
(3.43)
is obviously u = {1, 1}. Now, starting from u° = {0, 0}, let us apply (3.6), (3.7) to (3.43). Since u\ is the solution of Min (jv\ + \v11 — uO. we have u\ = 0. In a similar way we have u\ = 0, so that u" = {0, 0},
V n.
From the above result it follows that the algorithm (3.6), (3.7) does not converge to u. 3.6.2. Relaxation and nonfactorable convex sets In this section we assume that K is a closed convex subset of UN, such that N
If all the other assumptions of Sec. 3.1 hold, we can generalize algorithm (3.6), (3.7) as follows: u° e K, +
(3.44)! +
J(u\ \..., i d » r \ u?+ ! , . . . ) < J{u\ \..., utl
vt, M?+ 1 ; ...),
V{ti"1+1,...,«?+11,t>i,H7+1)...}eK,
{<'
«f+1,ai"+1)...}6K;
i = l,...,N.
(3.44)2
Since M° e K, it is very easy to prove that (3.44) is a well-posed problem, V n > 0. A very simple example will show that (3.44)!, (3.44)2 generally does not converge if K is nonfactorable. We take J = Jo denned by J(v) = v\ + v\ and K defined (see Fig. 3.1) by K = {v\v e U2, Vi > 0 , v2 > 0 , v,_ + v2> 1 } .
150
V Relaxation Methods and Applications Figure 3.1
The solution of
J(u) = MinJ(v)
(3.45)
veK
is obviously u = {%, %}. Then, starting from u° = {0, 1}, let us apply the algorithm (3.44)1;(3.44)2 to (3.45). It is easy to see that u" = {0,1} # u,V« > 0. Therefore we do not have convergence to the solution. Remark 3.8. Let's consider the "one-dimensional" torsion problem (3.46) of Chapter II, Sec. 3.7.1. In the sequel we assume that / e C ° [ 0 , 1]. Using the notation vh(xt) = vt, the above problem is approximated by the following variant of (3.41) (Chapter II, Sec. 3.7.1):
fduh/dvh_du, j0 dx \dx
uheKh,
dx
dx
where ft = f(xt),
(3.46)
i=l,...,N.
Problem (3.46) is equivalent to the minimization problem uheKh,
(3-47)
where ,v,
= {v J1 (fl v
h\ h)
\
OiV+1
h 2
N-l
Vi+1
= vN
- v,
N-l
-h^fiVt,
h
i=0
-o,
(3.48)
2
vi+i
-
h
Vi
£1,V/-*..
(3.49)
N-ll (3.50)
K,, is a nonfactorable closed convex subset of UN+1. However, it is proved in G.L.T. [1, Chapter 3], [3, Chapter 3] that (3.44),, (3.44)2 converges to the solution uh of (3.47) if/ > 0 on [0, 1] and if we start with u°h
4. Block Relaxation Methods
151
Since / > 0 implies M, > 0, V i = 1 , . . . , JV — 1, an obvious choice for ti° is u°h = 0.
4. Block Relaxation Methods In this section we take V = UN such that
V = f[Vt with V,= UNi,
(4.1)
where f; > 1 and
£ iV; = N-
(4-2)
If i> e F, then v = {vu ..., vq}, vt e Vt. We assume that K = f|9_
Kt, where
K; is a closed convex subset of Vt.
(4.3)
Then (4.3) implies that K is a closed convex subset of V. Finally we consider a convex functional J: V -> IR such that D(J) n X / 0 ,
lim
J(u) = + oo, J = J0 + Jl5
(4.4)
where J o G CX{V),
Jo strictly convex
(4.5)
and ,
(4.6)
where t h e ^ are convex, proper, and l.s.c. on V{. From Theorem 2.1 of Sec. 2 it follows that J(u) < J(v),
VoeK,
ueK,
(4.7)
has a unique solution (in D(J) n K) characterized by (J'0(U), » - u) + t 0"J(«I) - ./<(«<)) > 0,
V » 6 K,
ueK.
(4.8)
The generalization of the point relaxation algorithm (3.6), (3.7) is obviously u° e K, +
(4.9)
J « S . . . , Ultl u\ \ U"i+ ! , . . . ) < J ( M f \ . . . , «? + /, «,-, M?+l5 . . .), VvteKt, «-+1eXi; i = l , . . . , 9 . (4.10)
152
V Relaxation Methods and Applications
Algorithm (4.9), (4.10) is a block-relaxation algorithm. Using a variant of the proof of Theorem 3.1, it may easily be proved that the sequence u" defined by (4.9), (4.10) converges to the unique solution of (4.2). Remark 4.1. Most of the remarks we have made on algorithm (3.6), (3.7) still apply to (4.9), (4.10). Unfortunately, Remark 3.1 does not apply in general. Remark 4.2. In Cea and Glowinski [1] and Glowinski [6], it is proved that (4.9), (4.10) could also be used in infinite-dimensional situations.5 In this case the assumptions Jo e C 1 and strictly convex are not sufficient to insure the convergence of the relaxation algorithm. The basic reason for this is that closed bounded sets are generally noncompact in Banach spaces of infinite dimension. Using the same notation, in the two references cited above, it is proved that sufficient convergence conditions are J o is C 1 with J'o Lipschitz continuous on the bounded sets of V;
(4.11)
Jo locally uniformly convex, i.e., <J'0(i>) - J'M,v
- "> > K(\\v - «ll),
V u, v,
\\u\\ < M,
\\v\\ < M, (4.12)
where SM is the same as in Sec. 3.3.
5. Constrained Minimization of Quadratic Functionals in Hilbert Spaces by Under and Over-Relaxation Methods: Application We follow Cea and Glowinski [1, Sec. 3]. 5.1. Statement of the minimization problem Let V = Y\f= i Vt, where Vt is a real Hilbert space, V i = l,...,N. Norm and scalar product on Vi are, respectively, denoted by || • ||; and ((•, •));• If v e V, then v=
{vu...,vN}
with VteVi,
i=l,...,N.
On V we define a scalar product and a norm by
((u,v))=
£««*.»«))«, N
' More precisely, in reflexive Banach spaces.
\l/2
(5.1)
5. Constrained Minimization of Quadratic Functionals in Hilbert Spaces
153
V is a Hilbert space for ((•, •)) and || • ||. Let K be a closed convex subset of V such that K=
\[Ki,Ki±0,
Vi = 1
JV,
where Kt is closed and convex in Vt. Let J: V -> U be defined by J(v) = $a(v, v) - ((/, v)\
(5.3)
where the bilinear form a(-, •) on V x V, is continuous, symmetrical, and F-elliptic, i.e., 3 a > 0 such that a(u, u) > a|M|2,
V»eF,
and where feV. Under the assumptions on V, K, and J, the optimization problem J{u)
< J(v),
VOEK,
ueK
(5.4)
has a unique solution. This solution is characterized by a(u, v - u ) - ( ( / , v - u ) ) > 0 ,
V » e K,
u e K.
(5.5)
5.2. Some preliminary results From the properties of V it follows that
J(v) = J(vu ...,vN)
=\
X aifa, vj) - t <Wi, »d)i,
(5-6)
where the atj are bilinear and continuous on Vt x J^- with a y = a*. The forms flj,- are 1^-elliptic (with the same constant a). Using the Riesz representation theorem, it is easily proved that there exists Ay e Sf(Vj, V^ such that di/v;, Vj) = i(Vi, AijVj))t,
(5.7)
Atj = A%.
(5.8)
Moreover, the AH are self-adjoint and are isomorphisms from Vt to Vt. In the sequel it will be convenient to use the norm defined by au on Vt, i.e.,
Ilk-Ill? = (tufa, vd = ((.AtlVi,
vt))i,
i=l,...,N.
(5.9)
The norms || • |j ; and 111 • 111; are equivalent. The projection from Vt to Kt in the 111 • 111; norm will be denoted by P{. Before giving the description of the iterative method, we shall prove some basic results on projections, useful in the sequel. Let: (i) if be a real Hilbert space, with scalar product and norm denoted by (•, •) and || • ||, respectively, (ii) &(•,•) be a bilinear form on H, continuous, symmetric, and H-elliptic (i.e., 3 j8 > 0 such that b(v, v) > 0\\v\\2, V v e H).
154
V Relaxation Methods and Applications
Then from the Riesz representation theorem follows the existence of an isomorphism B: H -> H such that (Bu, v) = b(u, v),
V u,veH,
B = B*.
(5.10)
We denote by [-, •] and |-| the scalar product on H and the norm on H, respectively, denned by [w, D] = b(u, v),
Vu,veH,
(5.11)
\v\2 = b(v,v),
V»e/f.
(5.12)
The norms | • | and || • || are equivalent. Let (iii) C ^ 0 b e a closed convex subset of H and n be the projector from H ~* C in the | • | norm. (iv) j:H -» U be the functional defined by j(v) = $b(v,v)-(g,v),
Vcefl,
(5.13)
where g e H. Under the above assumptions, we have the following lemmas. Lemma 5.1. / / u is the unique solution of j(u)<j(v),
VveC,
usC,
(5.14)
then u = n(B-1g).
(5.15)
PROOF. The solution u of (5.14) is characterized by {Bu - g, v - u) > 0,
VceC, ueC.
(5.16)
From (5.16) it follows that (B(v-u), B-1g-u) = b(v-u,B-1g-u)<0,
VceC, u e C,
1
and (5.17) characterizes u as the projection on C of B~ g in the norm | • |.
(5.17) •
Lemma 5.2. Let uoe C and let u1 be defined by Uj = n(u0 + (a{B'lg - u0)),
co > 0.
(5.18)
Then 7(«o) - X « i ) >
2
-^
l«o — «i I2-
(5-19)
5. Constrained Minimization of Quadratic Functionals in Hilbert Spaces
155
PROOF. We have, V vu v2 e H,
Xvi) -J(P2) = Kloi - B" : 9l 2 - \v2 ~ B-W). i
(5.20)
1
Since uy — B~ g = u0 — B~ g + ux — u0, we have K - B~lg\2 = |«, - B - ^ l 2 + 2[H 0 - B " V « 0 - « J - |u 0 - u t | 2 .
(5.21)
From (5.18), and since u0 e C, it follows that
[w0 + a(B~lg - u0) - uu u0 - u{\ < 0. This implies ux\2 < co[u0 - B'ig,
\u0-
u0 - « J .
(5.22)
Then (5.19) clearly follows from (5.20)-(5.22).
5.3. Description of the algorithm For i = l,...,N, let the co, be positive numbers. Now let us consider the following algorithm: u° = {u°,...,
«$} arbitrarily given in K;
n+l
u" being known, we compute u
(5.23)
by
(5.24) Remark 5.1. It follows from (5.24) that the computation of u"+i can be achieved in three steps: Step 1. On Vt we minimize the functional
Hence we obtain a solution ,,"+1/3 _
d-l/f _ \
Step 2. We compute
M?+2/3
V J ,.»+! j
by
Step 3. At last we obtain u"+ i by We remark that if o); = 1, V i = l,...,N, then algorithm (5.23), (5.24) reduces to the block algorithm (4.9), (4.10) (see Remark 4.2).
156
V Relaxation Methods and Applications
5.4. Convergence of algorithm (5.23), (5.24) Proposition 5.1. We have J(u") - J(un+1)
> £ — - — - \\\u"+1 - M"|||,2-
(5.25)
PROOF. We have N
J{u") - J(un+")
= £ {J(u\+ \ . . . , uT-\, u i . . . . ) - J{u\+',...,
u1+ \ u " i + j , . . . ) ) .
(5.26)
Then (5.25) clearly follows from (5.24), and from the application of Lemma 5.2 at each of the differences of the right-hand side of (5.26). Proposition 5.2. IfOi<2,\/i=
I,... ,N,we have
J(un)>J(un+1),
Vn
and lim (un -un+1)
= 0 strongly in V.
(5.27)
PROOF. Since 0 < a>; < 2 implies (2 - cod/a, > 0, V i = 1 , . . . , N, it follows from (5.25) that J(u") is a decreasing sequence. Since J(u") is bounded below by J(u), where u is the solution of (5.4), J(u") is convergent. This implies lim (J(u") - J(u"+1)) = 0.
(5.28)
n-* + oo
Then, from (5.25), (5.27) and from (2 - a^/co; > 0, V i = 1 , . . . , N, it clearly follows that lim \\u1+1 - ^ | |
j =
0,
V J = 1,...,JV.
This implies (5.27).
From these two propositions we deduce: Theorem 5.1. / / 0 < cot < 2, V i = 1 , . . . , N, then the sequence u" defined by (5.23), (5.24) satisfies lim u" = u, n-* + co
where it is the solution of (5.4). PROOF. The F-ellipticity of a(-, •) implies a(u"+i
- u,u"+l
-u)>
a | | « " + 1 - u\\2.
(5.29)
From (5.29) it follows that a(u"+\ M" +1 - u) - ((/, u" + 1 - «)) > a{u, un+1 - u) - ((/, un+1 - «)) + a||U"+1 + u\\2. (5.30)
5. Constrained Minimization of Quadratic Functionals in Hilbert Spaces
157
Since u is the solution of (5.4), and since u"+i e K, we have a(u, u"+1 - w) - ((/, w"+1 - «)) > 0, which, combined with (5.30), implies a(un+l, u"+1 - u) - ((/, u" +1 - u)) > a\\un+1 - u\\2.
(5.31)
The left-hand side of (5.31) could be written as follows:
a(u"+\ u"+> - ii) - ((/, «" +1 - «)) = I ff I ^««J + 1 - / „ «? +1 - «,)) • (5.32) Let u"+1 be the vector of Kt for which the functional
attains its minimum on Kt. From Lemma 5.1 it follows that u"+l —PI A Mj
—
'I
1
'•
l f — V A u " + ' — V A ii" I I \J'
L*
'jui
Z^ Aijuj
(S 331 vJiJJ^
I r
Moreover, from the usual characterization of the minimum we have (AH01+1
+ I AijU"j + l + X A,tf - ft,v, - uTl))
> 0,
V o, 6 X,.
(5.34)
It follows from (5.32) that a(u"+\ un+1 - u) - ((/, u"+i - u))
(
i=l\\j
j>i
Since u, e Kt, (5.34) implies that the last term on the right-hand side of (5.35) is <0. Therefore (5.31), (5.35) imply
i=l\\j
( « ! i=l\\j>i
/
))
(5.36)
158
V Relaxation Methods and Applications
From (5.36) it follows that to prove the strong convergence of u" to u, it suffices to prove that the left-hand side of (5.36) converges to zero. Since the A{j are linear and continuous, the u1,u" are bounded uniformly in i and n, and limn^ + 00 \\u"+1 — u"\\ = 0, it follows from (5.36) that linv, + 0O ||fi" - u"\\ = 0 (where u" = {u"u ..., Wh ..., u"N}) will imply convergence. Let us prove the last property. Since a>; is >0, from (5.34) it follows that >uJ
))
(5.37)
From (5.24), (5.37), and since Pt is a contraction, it follows that
w s r ' - M r M i i ^ i i - ^ i i p r 1 -«?in,,
v »• = 1,2,..., AT.
(5.38)
Since 0 < Wj < 2 implies 0 < 11 - ct», | < 1, it follows from (5.38) that | | | t ? + 1 — wf"1"1 III,- < H i s r 1 - « ? l l l i .
Vi=l,...,iV.
(5.39)
F r o m the triangle inequality and (5.38), (5.39), it follows that III"? - n? + 1 |lli > III"? - S7 + I llli - lll"? + 1 - «? + 1 llli > ( 1 -ll-ffl.DIHfif+l-M-lll, > ( 1 _ | l _ < B | | ) | | | f i J + 1 - j i ? + 1 ||| l >
(5.40)
where 0 < 1 — 11 —
Remark 5.2. The above theorem generalizes Cryer [2], and also generalizes a classical result in finite dimensions (without constraints) for which we refer to R. S. Varga [1] and D. M. Young [1]. Remark 5.3. If to,- > 1 (resp., to,- = 1, co,- < 1), V i = 1 , . . . , N, then algorithm (5.23), (5.24) is an over-relaxation algorithm (resp., relaxation, underrelaxation algorithm) with projection.
5.5. Application to an elastic-plastic torsion problem 5.5.1. Statement of the continuous problem We consider the elastic-plastic torsion problem described in Chapter II, Sec. 3 (we use the notation of Chapter II, Sec. 3): a(u, v -u)>C
Jn Jn
(v -u)dx, - u) dx,
VceX,
where a(u, v) = | n VM • Vt> dx, and where K = {v\v e H},(Q), \Vv(x)\ < 1 a.e.}.
we K,
(5.41)
5. Constrained Minimization of Quadratic Functional in Hilbert Spaces
159
From the equivalence result of Brezis and Sibony [1] mentioned in Chapter II, Sec. 3.4, it follows that the solution of (5.41) is also the solution of a{u, v - u) > C \ (v - u) dx, Jn ue£={v\ve
Vcel,
H £(Q), | v(x) | < d(x, F) a.e.}.
(5.42)
In this section we consider only the numerical analysis of (5.42). 5.5.2. A finite element approximation of (5.42) We consider piecewise linear approximations only. As mentioned in Chapter II, Sec. 3.4, (5.42) is a variant of the obstacle problem we considered in Chapter II, Sec. 2. We assume that Q is a polygonal domain and that 2Th is a standard triangulation of Q (see Chapter II). From Chapter II, Sec. 2 it follows that Hj(Q) and K may be, approximated respectively, by6 Vh = {vh\vh e C°(Q), vh = 0 on F, vh\T e Pu V
Te.rh},
Kh = {vh\vh e Vh, \vh(x)\ < d(x, F), V x e ±h}.
Then (5.42) is approximated by a(uh, vh-uh)>C
(vh - uh) dx,
V vh e Kh,
uheKh.
(5.43)
Jn
From the results of Chapters I and II it easily follows that: Proposition 5.3. The approximate problem (5.43) has a unique solution. This solution is also the unique solution of the minimization problem J(uh)<J(vh),
VvheRh,
uheKh,
(5.44)
where J(v) = ja(v, v) — C \nv dx. Moreover, using the methods of Chapter II, we may prove the following: Theorem 5.2. Suppose the angles of ,Th are uniformly bounded below by 60>0ash-+0. Then lim ||uh - M|| H i (n) = 0,
where u and uh are, respectively, the solutions of (5.42) and (5.43). 6
t,h: set of the interior nodes of ^1.
V Relaxation Methods and Applications
160
Figure 5.1
5.5.3. Solution of the approximate problem The natural unknowns of the approximate problems (5.43) and (5.44) are uh(P), P e i j . Since vh -* J(vh) is quadratic with respect to vh(P), it follows from the structure of Kh that algorithm (5.23), (5.24) could be used to solve (5.43), (5.44). 5.5.4. Application to the case Q = ]0, 1[ x ]0, 1[ When Q = ]0, 1[ x ]0, 1[, or, more generally, a rectangle, it is convenient to use a finite element approximation which is actually equivalent to a finite difference approximation. Let AT be a positive integer and h = l/N. On Q we use the triangulation 2Th of Fig. 5.1. We have £„ = {My | My = {ihjh}, i,j integers, 0 < i,j < N],
By analogy with the finite difference notation, it is convenient to denote vh(Mij) by vu. Then the approximate problems (5.43), (5.44) are easily expressed in function of vu, since 2
Vlj-l
%1
\ _
h
2
/„
C
where vkl = 0 if Mkl $ £*, and since Kh = {vh\vh e Vh, \viS\ < du, 1 < i, j < N - 1}, where dtJ = d{Mtj, Y).
_
„
\2
5. Constrained Minimization of Quadratic Functionals in Hilbert Spaces
161
Algorithm (5.23), (5.24) has been used with cou = co, V 1 < ;', j < N - 1. Therefore the explicit form of (5.23), (5.24) is u° e Kh arbitrary given,
(5.45)
and for 1 < i,j < N - 1,
ui;1/2 = (1 - co)ulj + I (u1+ tJ + unij+! + « r h + U"J+JI + ^ O
( 5 - 46 )
«"/ l = max( — dij, min(w"/ 1/2 , di})),
(5.47)
where u£| = 0 if Mkl $ l,h. From Theorem 5.1 it follows that uh converges to the solution of (5.43), (5.44)ifO
I
K+1-M?;|<10-4.
(5.48)
The number of iterations necessary to obtain the convergence is given below as a function of co: CO
1
1.4
1.5
1.6
1.7
1.8
1.9
290
138
118
98
93
108
188
2
Number of
Iterations The C.P.U. time on an IBM 360/91 was 7.33 sec for co = 1 and 2.51 sec for co = 1.7. Figure 5.2 shows the behavior of
as a function of n (when co = 1.7). In the Fig. 5.3 we can see the elastic part of Q (in white) and the plastic parts of Q (striped area). In the plastic parts we have | VH | = 1 and | u(x) \ = d(x, F). In the elastic part we have — Aw = C and | u(x) \ < d(x, F). Remark 5.4. The optimal value of co is very close to the optimal value of co corresponding to the solution, by S.O.R., of the Dirichlet problem — Aw = C in the elastic part of Q (with the same discretization step h). This follows directly from the fact that the plastic part of Q 7 is very quickly obtained. From 7
It is the part of Q in which the constraints \u\ < d{x, T) are active (i.e., \u\ = d(x, F)).
V Relaxation Methods and Applications Figure 5.2
this fact it follows that the computation time is mainly used to solve — Au = C in the elastic part of £1 Moreover, the plastic part increases with C; this fact explains that the optimal value of a> and the number of iterations needed to obtain the convergence are, for a given h, decreasing functions of C (as is the computational time). Remark 5.5. In Cryer [1], [2] we can find a discussion on the choice of a>, when using point S.O.R. with projection to minimize quadratic functionals on a product of intervals.
Figure 5.3
6. Solution of Systems of Nonlinear Equations by Relaxation Methods
163
6. Solution of Systems of Nonlinear Equations by Relaxation Methods 6.1. Statement of the problem In this section we very briefly describe some relaxation methods for solving systems of nonlinear equation such that fi{ui,u2,...,uN)
= 0,
fi(u1,u2,.':.,uN)
= Q,
where ft: UN -> R. We define F: UN -+ UN by F = {fu...
(6.1)
,fN}.
6.2. A first algorithm We consider the following algorithm: M° given; n+1
u" being known, we compute u
(6.2)
by
m
-uX),
(6.3)
1 < i < N. If F = VJ, where J: UN -> U is a strictly convex C 1 function such that lim||i;||_ + 00 J(v) — +oo, then (6.3) has a unique solution (see Sec. 2). This solution is also the unique solution of J(u) < J(v),
Vc6R N ,
us UN.
(6.4)
Moreover, if co = 1, it follows from Theorem 3.1 that algorithm (6.2), (6.3) converges to the solution u of (6.1), (6.4). If F = VJ, co # 1, we refer to S. Schechter [1], [2], [3]. In these papers it is proved that under the hypothesis (i) J e C2(UN), (ii) (J'(w) - J'(v), w - v) > tx\\w - v\\2, V v, w; a > 0, (iii) 0 < co < coM, algorithm (6.2), (6.3) converges to the solution of (6.1), (6.4). Moreover, estimates of coM and of the optimal value of co are given. For the convergence of (6.2), (6.3) when F # VJ, we refer to Ortega and Rheinboldt [1], Miellou [1], [2], and the bibliography therein.
164
V Relaxation Methods and Applications
6.3. A second algorithm This algorithm is given by u° given; n+1
u" being known, we compute u
(6.5)
by
1 < i < N.
(6.6)
To our knowledge the convergence of (6.5), (6.6) for co ^ 1 and F nonlinear has not yet been considered. Remark 6.1. Algorithms (6.2), (6.3) and (6.5), (6.6) are identical if co = 1 and/or F is linear. Remark 6.2. In many applications, from the numerical experiments it appears that (6.5), (6.6) is faster than (6.2), (6.3), co having its (experimental) optimal value in both cases. Intuitively this seems to be related to the fact that (6.5), (6.6) is "more implicit" than (6.2), (6.3). For instance, (6.5), (6.6) could easily be used if F is only denned on a subset D on [R*; in such a situation, when using (6.2), (6.3) with co > 1, it could happen that {u"+ 1,...,u"+i, u"+ u .
6.4. A third algorithm In this section we assume that F e C^M"). A natural method for computing w" +1/2 in (6.3) or u"+l in (6.6) is Newton's method. We recall that Newton's method applied to the solution of the single-variable equation /(*) = 0 is basically: x° given;
(6.7) m
f(x }
4+1/2
(6-8)
+1
In the computation of u" in (6.3) or u" in (6.6) by (6.7), (6.8), the obvious starting value is u". Then obvious variants of (6.2), (6.3) and (6.5), (6.6) are obtained if we run only one Newton iteration. Actually, in such a case, (6.2), (6.3) and (6.5), (6.6) reduce to the same algorithm, which is u° given;
(6.9)
6. Solution of Systems of Nonlinear Equations by Relaxation Methods
165
In S. Schecter, loc. cit., the convergence of (6.9), (6.10) is proved, if F = VJ, under the same assumptions as in Sec. 6.2 for algorithm (6.2), (6.3) (with a different <x>M in general). Remark 6.3. In Glowinski and Marrocco [1], [2] and Concus [1], we can find comparisons between the above methods when applied to the numerical solution of the nonlinear elliptic equation modeling the magnetic state of ferromagnetic media (see also Winslow [1]). Applications of the first algorithm for solving minimal surface problems may be found in Jouron [1].
CHAPTER VI
Decomposition-Coordination Methods by Augmented Lagrangian: Applications1
1. Introduction 1.1. Motivation A large number of problems in mathematics, physics, mechanics, economics, etc. may be formulated as Min {F(Bv) + G(v)},
(P)
where • V, H are topological vector spaces, • F: ff -> R, G:V->U are convex proper l.s.c. functional. Let us give two examples taken from Chapter II. EXAMPLE 1 This is the Binghamflow problem of Chapter II, Sec. 6; we recall that with Q a bounded domain of U2, we consider the variational problem
Min 1^ f \Vv\2dx + g\ \Vv\dx - f fvdxl,
(1.1)
where v and g are positive constants. Then (1.1) is the particular problem (P) in which • • . . •
V = H J(Q), H = L2(Q) x L2(Q), £ = V, F(q) = {vl2)\n\q\2 dx + g \n\q\dx(\q\ G(«)= -\nfvdx.
Actually, we can also take • F(q) = gla\q\dx, • G(v) = 1
This chapter follows Fortin and Glowinski [2].
= Jq{ + q\),
1. Introduction
167
EXAMPLE 2 This is the elastic-plastic torsion problem of Chapter II, Sec. 3; with Q still bounded in U2, we consider
Minj^ f | V i ; | 2 ^ - f fvdxl,
(1.2)
where
K = {veHKCl Problem (1.2) is the particular problem (P) in which • V = tf £(fi), H = L2(Q) x L2(Q), . B = V, . G(t;) =
-\afvdx,
where /^ is the indicator functional of the convex set K={qeH, "0
M < 1 a.e.}, if qeK
We can also take = /*,
Synopsis. Problems of type (P) have a special structure, and in the sequel we shall introduce iterative methods of solution taking it into account. 1.2. Principle of the methods The decomposition-coordination methods to follow are based on the following obvious equivalence result. Theorem 1.1. (P) is equivalent to Min {F(q) + G(v)},
(n)
{vq}EW
where W = {{v, q} e V x H, Bv - q = 0}. In the sequel we shall assume that V and H are Hilbert spaces with inner products and norms denoted by ((•, •)), ||-|| and (•, •), \-\, respectively. We then define a Lagrangian functional .Sf associated with (n), by JSf (c, q, ft) = F(q) + G(v) + (p, Bv - q),
(1.3)
168
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
and for r > 0 an Augmented Lagrangian ^Cr is defined by Ser(v, q, n) = &(v, q,fi) +
r
-\Bv-q\2.
(1.4)
Remark 1.1. Augmented Lagrangian methods for solving general optimization problems have been introduced by Hestenes [1] and Powell [1]. Augmented Lagrangian methods for solving problems like (P) via (n) have been introduced by Glowinski and Marrocco [3] and also [4]-[6].
2. Properties of (P) and of the Saddle Points of <£ and <er 2.1. Existence and uniqueness properties for (P) Let us define J: V -> U by J(v) = F(Bv) + G(v). Then (P) can also be written J(u)<J(v),
\fveV, ueV.
(2.1)
Let j : X ->• U; we define the so-called domain ofj(-) by dom(j) ={xeX,
j(x) e R}.
Then, if dom(F o B) n dom(G) # 0 ,
(2.2)
/ is convex, proper, and l.s.c.. Therefore, sufficient conditions for (P) to have a unique solution are (cf. Cea [1], [2] and Ekeland and Temam [1]): |M| = +GO, • J strictly convex. Remark 2.1. If B is an injection from V to H, with R(B) ( = range of B) closed in H, then | J3i?| is a norm on V equivalent to \\v\\.
2.2. Properties of the saddle points of =£? and i£r We have: Theorem 2.1. Let {u, p, X} be a saddle point of $£ onV x H x H; then {u, p, X} is also a saddle point of J£r, V r > 0 and conversely. Moreover, u is a solution of (P) and p = Bu.
2. Properties of (P) and of the Saddle Points of if and <£,
169
PROOF, (i) {u, p, A} be a saddle point of if; then if(u, p, A) e IR and , p, /x) < i f (u, p, X) < ^{v, q, A), V {v, q, n) 6 V x H x tf, {«, p, A} e F x # x H.
(2.3)
From the first inequality (2.3) and from (1.3) it follows that (jt, Bu-p)<
(A, Bu-p),
V n 6 H,
which obviously implies that Bu = p.
(2.4)
From the second inequality (2.3) and from (1.3) and (2.4) it follows that J(u) = <£{u, p , X) < £C(v, q, X),
V{v,q}eV x H.
(2.5)
Taking q = Bv in (2.5), it follows from (1.3) that J(u)
< £C(v, Bv, X) = J(v),
V V E V ;
(2.6)
hence u is a solution of (P). Since p = Bu, we have <£r{u, p, n) = J?(u, p, n) = J(u),
MneH;
(2.7)
it then follows from (2.3), (2.7) that if r (u, p, n) = £?r(u, P, X) < &(v, q, X),
V {v, q, n) e V x H x H.
(2.8)
2
Since &,(», 9, fi) = ^f(", 1, n) + (r/2)\Bv - q\ , from (2.8) we obtain if r («, P, n) < yr(u, P, X) < £er(v, q, X),
V {v, q, n) e V x H x H,
(2.9)
which proves that {«, p, X} is also a saddle point of if,. onV x H x H.To conclude this part of the proof, we observe that from (2.3), {u, p} is a solution of i f (u, p , A) < &(v, q, X),
V {v, q}eVxH,
[u, p} e V x H,
(2.10)
from which it follows that {u, p} is characterized (see Cea [1], [2], Ekeland and Temam [1], and also Chapter V, Sec. 2) by F(q) ~ F(p) - (X, q - p) > 0, G(v) - G(u) + (X, B(v - «)) > 0,
V q e H, p e H,
(2.11)
V v e V, u e V.
(2.12)
(ii) Let {u, p, X] be a saddle point of ifr with r > 0. Then, as in part (i), this implies that p = Bu and that u is a solution of (P). Moreover, since {u, p, A} is solution of i f r ( u , p , X) <
V{v,q}eV
x H, {u, p} e V x H,
(2.13)
it is characterized by F(q) - F(p) + r(p - Bu,q - p) - (X,q - p)>0, G(v) - G(u) + r(Bu - p, B(v - «)) + (A, B(v - u)) > 0,
VqeH,
peH,
VueF,
ueV.
(2.14) (2.15)
But since Bu - p = 0, (2.14), (2.15) reduce to (2.11), (2.12), and this fact implies that {«, p, A} satisfies (2.10). It then follows from (2.7) that {u, p, A} satisfies (2.3), and this completes the proof of the theorem. •
170
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
3. Description of the Algorithms From Theorem 2.1 it follows that a method for solving (P) is to solve the saddle-point problem £er(u, P, H) < &r{u, p , X) < Ser{v, q, X), V {v, q, n} e V x H x H,
{u, p,X}eVxHxH.
(3.1)
To do this we shall use (see Cea [1], and G.L.T. [1, Chapter 2], [3, Chapter 2]) an algorithm of Uzawa type and a variant of it. 3.1. First algorithm We denote by ALG 1 the following algorithm: X°eH
given;
(3.2)
then, X" known, we define u", p", X"+' by jS?r(«"» V\ A") < <£r(v, q, X"),
V {v, q} e V x H, {«", p"}eVx
X»+1 = X" + pn(Bu" - p"),
Pn
> 0.
H, (3.3) (3.4)
Problem (3.3) is in fact equivalent to the following system of two coupled variational inequalities (of the second kind): G(v) - G(un) + (Xn, B(v - «")) + r(Bun - p", B(v - u")) > 0, \fveV, u"eV, (3.5) F(q) - F(p") - (Xn, q - p") + r(p" - Bit", q - p") > 0,
V q e H,
pneH. (3.6)
The convergence of ALG 1 will be studied in Sec. 4. 3.2. Second algorithm The main drawback of ALG 1 is that it requires the solution of the coupled EVI's (3.5), (3.6) at each iteration. To overcome this difficulty, it is natural to consider the following variant of ALG 1 (denoted ALG 2 in the following): {p°, Xl}eH
x Hgiven;
n
(3.7)
n+1
then, {p"~\ X"} known, we define {u , p", X } by G(v) - G(un) + (Xn, B(v - u")) + r(Bu" - pn~ \ B(v - w")) > 0, V v e V, u"e V, (3.8) n F(q) - F(p") - (X , q- p") + r(p" - Bit", q - p") > 0, V q £ H, p"eH, (3.9) +i
X"
n
= X" + pn(Bu -p"),
Pn>0.
The convergence of ALG 2 will be studied in Sec. 5.
(3.10)
4. Convergence of ALG 1
171
4. Convergence of ALG 1 4.1. General case In this subsection, V and H are possibly infinite dimensional; of course we assume that dom(F o B) n dom(G) # 0 ,
(4.1)
B is an injection and R(B) is closed in H.
(4.2)
and also
We also assume that /
lim I 4 I - + 00
^ ) = +oo,
(4.3)
\i\
F = F o + -f 1 with ^o? ^1 convex, proper, and l.s.c,
(4.4)
Fo is Gateaux-differentiable and uniformly convex on the bounded sets of H.
(4.5)
By definition we say that Fo is uniformly convex on the bounded sets of H if the following property holds: V M > 0, 3 SM: [0, 2M] -> U, continuous, strictly increasing with «5M(0) = 0, such that V q , p e H with \p\ < M, \q\ < M we have (F'0(q) - F'0(p), q - p) > 8M(\q - p \ ) , (4.6) where F'o = VF 0 is the G-derivative of Fo. From the above properties, (P) has a unique solution u, and we define p e H by p = Bu. EXERCISE
4.1. Prove that (P) is well posed if (4.1)-(4.5) hold.
On the convergence of ALG 1, we have: Theorem 4.1. We assume that =Sf has a saddle point {u, p, A} £ V x H x H. Then under the above assumptions on B, F, G and if 0 < a 0 < pn < «! < 2r,
(4.7)
the following convergence properties hold: u" -> it strongly in V,
(4.8)
p" -> p = Bu strongly in H,
(4.9)
n
X"+1 _ X ->• 0 stron^/y m H, n
X is bounded in H.
(4.10) (4.11)
172
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
Moreover, if X* is a weak cluster point of {X"}n in H, then {it, p, X*} is a saddle point of !£r over V x H x H. PROOF. Since {u, p, X} is a saddle point of <£,, we have ^r{u,p,X)<^r(v,q,X),
V{v,q}eVxH,
{u,p}eVxH.
(4.12)
Therefore {u, p} is characterized by G(v) - G(u) + (X, B(v - «)) + r(Bu - p, B(v - u)) > 0, (F'o(p),q-p) + Fl(q)-F1(p)-(lq-p)
V v e V, ueV,
+ r(p-Bu,q-p)>0,
(4.13)
MqeH, peH. (4.14)
Moreover, from Theorem 2.1 we have Bu = p; therefore X = I + Pn(Bu - p).
(4.15)
Let us define u", p", 1" by u" = u" - u,
p" = p" - p ,
i " = X" - I.
From (3.4), (4.15) it then follows that 1" + 1 = I" + Pn(Bu" - p"), which implies |!»+i| 2 = |A"|2 + 2pn(l\ Bu" - p") + p2n\Bu" - p"| 2 or, more conveniently, \l"\2 - | I " + 1 | 2 = -2pn(X",Bu" - p") - pl\Bu" - p"\2.
(4.16)
Since {u", p"} is a solution of (3.3), it is characterized by G(v) - G(u") + {X", B(v - un)) + r(Bu" - p", B(v - «")) > 0,
VceF,
u" e V, (4.17)
(F'0(p"), q-p") + F.iq) - Fj(p") - (X", q - p") + r(p" - Bu", q - p") > 0, V q e H, p" e H. (4.18) Taking v = u (resp., v = u") in (4.17) (resp., (4.13)) and q = p (resp., q = p") in (4.18) (resp., (4.14)), we obtain, by addition, {I", Bu") + r(Bu" - p", Bu") < 0,
(4.19)
(F'0(p") - F'0(p), p") - (I", p") + r(p" - Bu" • p") < 0,
(4.20)
which imply, also by addition, (1", Bu" - p") + (F'0(pn) - F 0 (p), p") + r\Bu" - p"\2 < 0, i.e., - ( 1 " , Bu" - p") > (F'0(p") - F'0(p\ pf) + r\Bu"-p"\2.
(4.21)
Combining (4.16) and (4.21), we obtain |i"| 2 - U " + ' I2 > 2pn{F'0(p") - F'0(p), p") + Pn{2r - pn)\Bu" - p"| 2 > 0.
(4.22)
4. Convergence of ALG 1
173
Assuming that (4.7) holds, it follows from (4.22) that lim I B S " - p " | = 0,
(4.23)
n-> + oo
lim (F'0(p") - F'Q(p), p"-p)
= 0,
(4.24)
n-» + oo
A" is bounded in H.
(4.25)
Since p = Bu, it follows from (4.23) that lim \Bu" - p"\ = 0.
(4.26)
n-> + oo
Since F is proper, there exists poe H such that F(p0) e R; then, from the characterization (3.6), we have F(Po) - (*", Po) + r(p" - Bu", p 0 ) > F(p") - (A", p") + r(p" - Bu\ p").
(4.27)
Since A" and p" — £«" are bounded, (4.27) implies A) > F(p") - Pi\pn\,
(4.28)
where jS0, /?i are independent of n. From (4.3), (4.28) it then follows that p" is bounded in H, i.e., 3 M such that | p" | < M,
V «.
(4.29)
Then, using the uniform convexity property (4.5), (4.6) of F o , we obtain from (4.29) (assuming M > \p\) that ( W ) - F'0{p), f ~ p ) > 5M(\p"
- p\\
which combined with (4.24), implies, lim <5M(IP" - p|) = O o lim \p" - p\ = 0.
(4.30)
n~* + X
It then follows from (4.26), (4.30) that lim Bu" = p = Bu strongly in H.
(4.31)
n-* + oo
Since B is an injection with R(B) closed in H, then (4.31) implies that lim u" = u strongly in V.
(4.32)
n-* + oo
The convergence result (4.10) clearly follows from (4.7), (4.26). Let X* be a weak cluster point of {/!"}„ in H. Then, passing to the limit in (3.5), (3.6) and using the l.s.c. property of F and G, we have G(v) + (A*, B(v - u)) + r(Bu - p, B(v - u)) > lim inf G(u") > G(u), F(q) - (A*, q-p)
+ r(p-Bu,q-p)>
lim inf F(p") > F(p),
V v e V, VqeH,
ueV,
peH,
i.e., G{v) - G(u) + (A*, B(v - u)) + r(Bu - p, B{y - «)) > 0, F(q) - F(p) - (A*, q-p)
+ r(p-Bu,q-p)>Q,
VveV,
u e V, (4.33)
V q eH, peH.
(4.34)
174
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
As noted before (see (2.13)-(2.15)), (4.33), (4.34) is equivalent to &r(u,
p , k*) < &r(v, q, A*),
V {v, q} e V x H,
{ u , p}eV
x H.
(4.35)
Since, from p = Bu, we have ££r(u, p, jj) = £?{u, p, ix) = J(u),
V fieH,
we obtain Ser{u, p , fi) = £er(u, p , A*),
V fi e H.
(4.36) 2
It clearly follows from (4.35), (4.36) that {u, p, X*} is a saddle point of £Cr; this completes the proof of the theorem. •
4.2. Finite dimensional case If V and H are finite dimensional, we have convergence of ALG 1 with weaker assumptions on F, B, and G than in Sec. 4.1. The reasons for this are the following: (1) Since the constraint Bv — q = 0 is linear, if (P) has a solution, then if and j§^ have a saddle point (see Rockafellar [1] and Cea [1], [2]). (2) R(B) is always closed. (3) It follows from Cea and Glowinski [1] (and from Chapter V, Sec. 3.3) that Fo satisfies the uniform convexity property (4.5), (4.6) if Fo is C 1 and strictly convex. (4) If Fo is C 1 and strictly convex, then F'o is C° and strictly monotone, i.e., (F0(12) -
F
'o(iiX I2 - 0,
V qu q2 e H,
q1 /
q2.
Then if (P) has a solution, the property
lim ^ f = + ° o is not necessary. This is related to the following: Lemma4.1. Let H be finite dimensional and A:H^H be continuous and strictly monotone. Let {#"}„>0, Pn e H,V n, and p e H be such that lim (A(p") - A(p), p" - p) = 0;
(4.37)
H-» + 00
then lim f = p.
(4.38)
n~* + 00
2 Concerning the convergence of {X"}a, we have, in fact, a weak convergence of the whole sequence to a X* such that {u, p, A*} is a saddle point of i£r (see Sec. 7).
4. Convergence of ALG 1
175 Figure 4.1
PROOF. Assume that (4.38) does not hold. Then there exist 5 > 0 and a subsequence, denoted {pm}m, such that \pm-p\>6, Let S(p; 5/2) = {qeH,\q-p\
Vm.
(4.39)
= 6/2}. We define zm e S(p; 6/2) by 5 pm — p Z P
+
^
m
(440)
2^J\;
m
then z e ]p, p \_ a H (see Fig. 4.1). We introduce tm = 6/2\pm - p\; then zm = p + tm(pm - p),
(4.41)
0 < tm < \.
(4.42)
and from (4.39),
Since A is strictly monotone, we have {A{f) - A(p), pm - p) > (A(p + t(pm - p)) - A(p), pm - p),
Vt6]O, 1[; (4.43)
m
then, taking t = t in (4.43), we obtain (A(pm) - A(p), pm-p)>
(A{zm) - A(p), pm-p)>
0.
(4.44)
From (4.41), (4.42), and (4.44), it then follows that (A(pm) - A(p), Pm-P)>^,
(A(zm) - A(p), zm-p)>
2(A(zm) - A(p), z™ - p)
> (A(zm) - A(p), zm - p)> 0.
(4.45)
Since S(p, 8/2) is compact, we can extract from {zm}m a subsequence—still denoted {zm}m—such that lim zm = z,
z e s(p, - ) .
(4.46)
Since A is continuous, it follows from (4.37), (4.45), and (4.46) that (A(z) - A(p), z - p) = 0.
(4.47)
176
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
The strict monotonicity of A and (4.47) imply that z = p, which is impossible since \p — z\ = 6/2. Therefore (4.39) cannot hold => limn^ + 0O p" = p. •
From the above properties we can easily prove the following: Theorem 4.2. Assume that V and H are finite dimensional and that (P) has a solution u. We suppose that: • B is an injection, • G is convex, proper, and l.s.c, • F = Fo + Ft with Fj convex, proper, and l.s.c. over H and Fo strictly convex and C1 over H. Then (P) has a unique solution, and if 0 < a 0 < pn < ati < 2r holds, we have, for ALG 1, the following convergence properties: lim u" = u, -* + 00
lim p" = Bu, -> + oo
lim X"+ i
_ ;
X" is bounded in H. Moreover, if X is a cluster point of {/l"}n, then {u, p, X} is a saddle point of 5£r over V x H x H.
4.3. Comments on the use of ALG 1. Further remarks Assume that r is fixed and that we use a fixed value p for pn. Then, from our computational experience, it appears that the best convergence is obtained for p = r. Concerning the choice of r, it can be theoretically proved that the larger the r, the faster the convergence; practically, the situation is not so simple, for the following reasons. The larger the r, the worse the conditioning of the optimization problem (3.3) (or of the equivalent system (3.5), (3.6)). Then, since (3.3) is numerically (and not exactly) solved, at each iteration an error is made in the determination of {«", p"}. The analysis of this error and the effect of it on the global behavior of ALG 1 is a very complicated problem since we have to take into account the conditioning of (3.3), the stopping criterion of the algorithms (usually iterative) solving (3.3), roundoff errors, etc. Fortunately, it seems that the combined effect of all these factors is an algorithm which is not very sensitive to the choice of r (see Glowinski and Marrocco [5] and Fortin and Glowinski [1] for more details).
4. Convergence of ALG 1
177
From a numerical point of view, the only nontrivial part in the use of ALG 1 is the solution at each iteration of problem (3.3). Taking into account the particular structure of (3.3), it follows from Chapter V, Cea and Glowinski [1], and Cea [2] that a method very well suited to the solution of (3.3) is the block-relaxation method described below: All the problems (3.3) are of the following type: J2?,(K, p , H) < Ser{v,
q, n),
V {», q} e V x H , {u, p } e V x H , (4.48)
where \i is given. The minimization problem (4.48) is equivalent to the system
G(v) - G(u) + (n,B(v - u)) + r(Bu - p,B(v - u)) > 0,
VveV,ueV, (4.49)
F{q) - F(p) - {ii, q - p) + r(p - Bu, q - p) > 0,
V«-efl,
p e H. (4.50)
Then a block-relaxation method for solving (4.49), (4.50) is {u°,p0} given;
(4.51)
then, {um, pm} known, we obtain {um+ \ pm+1} from G(v) - G(« m+1 ) + (/i, B(v - wm+1)) + r(Bum+1 - pm, B(v - um+ x)) > 0, m+i
F(q) - F(j>
) -(n,q-
m+1
p
) + r(p
m+1
\fve V, um+1e V, m+1
- Bu
(4.52)
m+1
, q - p ) > 0, pm+1eH.
V<JEJFJ,
(4.53)
Sufficient conditions for the convergence of (4.51)-(4.53) may be found in Chapter V, Cea and Glowinski, and Cea, loc cit. In practice, when using (4.51)-(4.53), a stopping test of the following type will be used: Max(|| M m+1 - um\\, \pm+1 - pm\) < e. Another possibility is to stop after a fixed number of iterations. If, for instance, we stop after only one iteration of (4.51)-(4.53), and if at iteration n of ALG 1 we initialize with {M"" 1 , p"" 1 } the computation of {np, p"} by (4.51)-(4.53), then we recover ALG 2. Remark 4.1. Other relaxation methods can also be used; moreover, it can be worthwhile to introduce over-relaxation parameters to increase the speed of convergence of (4.51)-(4.53). Remark 4.2. The choice p = r may be motivated by the following proposition.
178
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
Proposition 4.1. Suppose that F(q) = j\q\2 and that G is linear. Then V 1° e H, we have, for the sequence {«"}„ of A L G 1, convergence to the solution u of (P) in less than three iterations if we use pn = p = r,r given. Preliminary Remark. In the above situation we have (P) equivalent to B'Bu = f,
(4.54)
where G(v) = ((/, v)), V v e V. Therefore, using ALG 1 for solving (P) has no practical interest. But even in this trivial case we shall observe that the behavior of ALG 1 is "interesting" since the convergence of u" in a finite number of iterations does not imply a similar convergence property for p" and X". PROOF OF PROPOSITION 4.1 From (4.17), (4.18) it follows that in the particular case we are considering, ALG 1 reduces to
X° given in H;
(4.55)
rB'Bu" = rB'p" - B'X" + / ,
(4.56)
n
p" = X + r(Bu" - p"), A"
+1
(4.57)
= X" + r{Bu" - p").
(4.58)
We can easily prove that the unique saddle point of i f r over V x H x H is {u, Bu, Bu}, i.e., p = Bu, X = Bu; using the notation u" = u" — u, p" = p" — p, X" = A" — X, it follows from (4.56)^(4.58) that B'Xn + rB\Bu" - p") = 0, +1
n
X"
+
= p => X" * = p",
p" = X" + r(Bu" - p"),
V n > 0,
(4.59)
V n > 0,
(4.60)
V n > 0.
(4.61)
Multiplying (4.61) by B' and comparing with (4.59), we obtain B'p" = 0,
V n > 0.
(4.62)
Since (4.60), (4.61) imply pn+1 =p«
+ r(BU-+1
_ p»+i);
v«>0,
(4.63)
multiplying by B' and taking (4.62) into account, we obtain B'Bu"+ x = 0,
V n > 0 => Bu"+1 = 0,
V n > 0.
(4.64)
Since \Bv\ is a norm on V, (4.64) implies that u" = u, V n > 1. Hence the convergence of u" to u requires two iterations at most. Using (4.63), (4.64), we have "+'
1+r
p",
V n > 0.
(4.65)
From (4.65) it follows that the larger the r, the faster p" converges to p = Bu; for more details on the convergence of p", see Fortin and Glowinski [2]. •
5. Convergence of ALG 2
179
5. Convergence of ALG 2 5.1.
Synopsis
In this section we shall prove that under fairly general assumptions on F and G, we have convergence of ALG 2 if 0 < pn = p < [(1 + -v/5)/2]r. We do not know if this result is optimal, since in some cases (G linear, for example) the upper bound of the interval of convergence is 2r. Actually this question is rather academic, since in the various experiments we have done with ALG 2, the optimal choice seems to be p = r.
5.2. General case We study the convergence of ALG 2 with the same hypotheses on B, F, G as in Sec. 4.1. We then have: Theorem 5.1. We suppose that ^£r has a saddle point {u, p, X} over V x H x H. Then, if the assumptions on B, F, G are those of Sec. 4.1 and if
0
(5.1)
we have the following convergence properties: u" —> u strongly in V,
(5.2)
p" -*• p strongly in H,
(5.3)
X«+1 _ X« _• 0 strongly
in H,
(5.4)
X" is bounded in H.
(5.5)
Moreover, if X* is a weak cluster point of {X"}n, then {u, p, X*} is a saddle point of£eroverV x H x H. PROOF.
Let us again define u", p", 1" by u" = u" - u,
p" = p" - p,
I" = X" - X.
Since {u, p, X} is a saddle point of i£r over V x H x H, we have G(v) - G(u) + (X, B(v - u)) + r(Bu - p, B(v - «)) > 0,
V v e V,
(F'0(p), q - p) + Ft(q) - F\(p) - (X, q - p) + r(p - Bu, q - p) > 0 X = X + p(Bu - p).
(5.6)
V q e H, (5.7) (5.8)
180
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
Moreover, (3.8)-(3.10) imply G(i>)- G(u") + (An, B(v - u")) + r(Bu" -p"-\B(v
- « " ) ) > 0,
V v e V,
( W X 9 - P") + F^q) - F.ip") - (A", q - f) + r(p" - Bun, q - p") > 0, V q e H, +1
r
(5.9) (5.10)
= r + P(Bu"-p").
(5.ii)
Taking v = u" (resp. t> = u) in (5.6) (resp., (5.9)) and q = p" (resp., q = p) in (5.7) (resp., (5.10)), we obtain, by addition, r(Bif - pn'\ Bun) + (1", BM") < 0,
(5.12)
(F'oipl ~ F'0(p), p") + r(p" - 5«", p") - (1", p") < 0.
(5.13)
By addition of (5.12) and (5.13), it follows that (F'0(p") - F'0(p), p"-p) + r\Bu" - p"\2 + (I", Bu" - p") + r(p" - p"~ \ Bit") < 0. (5.14) By subtracting (5.8) from (5.11), we obtain |!"| 2 - | I " + 1 | 2 = -2p(Bu" - p",X") - p2\Bu" - p"| 2 .
(5.15)
From (5.14), (5.15) it then follows that | A f - \X"+l\2 > 2p(F'0(p") - F'Q(p),p") + p(2r - p)\Bu" - p"\2 + 2pr(p" -
p"-\Bu"). (5.16)
Starting from Bu" = (Bu" - Bunl)
+ (Bu"~s - p"" 1 ) + p " - \
we obtain (Bu", pn - p"~') = (Bu" - Bu"~l,p" - p"~l) + (Bu"~l -p"-\p"
~ p"-1) + (p"-1,p"-p"-1).
(5.17)
Since (p"- 1 , p" - ff-1) = i ( | p f - Ip"-112 - \p" - p""112), it follows from (5.17) that 2pr(Bu", p" - p " " 1 ) = 2pr(Bu" - Bunl, p" - p " " 1 ) + 2pr(Bu"~1 + pr(\p"\2-\p"-1\2-\p"-r-1\2).
- p " ~ \ p" - p " " 1 ) (5.18)
Taking (5.10) at n — 1 instead of n, we have (F'0(p"- % q-p"~1) + FM - F^"^) + r(p"~1 - Bu"' \ q - f~l) > 0.
-(X"~\q-
p"~l) (5.19)
Taking q = p"~1 in (5.10) and q = p" in (5.19), we obtain, by addition, (F'0(p") - F 0 (p"-'), p" - p""') - (ln - l"~\ p" - p"-1) + r\p" - p-112 -r(Bu" - Bu"~ \ p" - p"- 1 ) < 0. (5.20)
5. Convergence of ALG 2
181
But since F'o is monotone, it follows from (5.20) that r\F ~ F~112 - (^ - I-"' \ p" - f-1)
- r(Bu" - Bu"~\ p" - F'1) < 0.
(5.21)
We have (from (3.10)) X" = A"" 1 + p(Bu"~l
-
p"~l),
1" - l"~1 = p(Bu"-'
- p"'1).
which implies that (5.22)
From (5.21), (5.22) it then follows that r\p" - p"" 1 ! 2 - p{Bu"-1 - p"~\ p" - p"'1) - r{Bu" - Bu"'1^"
- p"" 1 ) < 0,
i.e., r{Bu" - Bu"~ \ p" - p"'1)
p"'1 \2 - piBu"'1
>r\p"~
- p"~ \ p" - p " ~ ! ) .
(5.23)
From (5.18), (5.23) it then follows that 2pr(Bu",p" - p"-1) > pr(\p"\2 - |p"^| 2 ) + pr\P" -
F'1?
+ 2p(r - p)(Bu"-1 - p"-\ p" - p"~l).
(5.24)
Finally, combining (5.16) and (5.24), we obtain (\l"\2 + pr|p""M 2 ) - (U" + 1 | 2 + pr\T\2) > MF'O(P") - F'Q(p),jr) + p(2r - p)\Bu" - pn\2 + pr\p" - p"'1]2) + 2p(r - p)(Bu"-1 - pn~ \ p" - f ~ ' ) .
(5.25)
Then using, the Schwarz inequality, it follows from (5.25) that, V a > 0, we have (|Af + H p " " T ) - (U" +1 | 2 + pr\p"\2) > 2p(F'0(p") - F'0(p),p") + p(2r - p)\Bu" - p"\2 + pr\? - p"~'\2 - p\r - PlHBu"-1
-F~l\2
+ *\F-r-1\2\
(5-26)
If p = r, it is clear that, using the same method as in the proof of Theorem 4.1, we have (5.2)-(5.5). If 0 < p < r, taking a. = 1 and observing that | r — p | = r — p, from (5.26) we observe that (\X"\2 + pr\p-l\2 - (|1"
+1 2
|
+ p(r - p^BO-1
+ H p f
- p"" 1 ! 2 )
+ P(r - P)\BU" -
F\2)
n
> 2p(F>0(p") - F'0(p), F) + pr\Bu - p"| 2 + p 2 |p" - p"" 1 1 2 > 0, which clearly implies (5.2)-(5.5). If p > r, we have \r — p\ — p — r, and then it follows from (5.26) that (5.2)-(5.5) holds if we have p < pM, where - pM) = - PM(PM
~r),
a ~ r). By elimination of a, it follows from (5.27) that PM
- ?PM - r2 = 0,
(5.27)
182
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
i.e. (since pu > 0),
_ 1 + V5 Pu —
2
r
'
Then, using basically the same method as in the proof of Theorem 4.1, we can easily prove, from (5.2)-(5.5), that {u, Bu, X*} is a saddle point of !£r over V x H x H if X* is a weak cluster point of {X"}n. •
5.3. Finite dimensional case Using a variant of the proof of Theorem 5.1, and Lemma 4.1, we can easily prove: Theorem 5.2. Assume that the assumptions on V, H, F, B, G are those of the statement of Theorem 4.2. Then if
the conclusions of the statement of Theorem 4.2 still hold.
5.4. Comments on the choice of p and r 5.4.1. Some remarks Remark 5.1. If G is linear, it has been proved by Gabay and Mercier [1] that ALG 2 converges if 0 < pn = P < 2r. The proof of this result is rather technical, and an open question is to decide if it can be extended to the more general cases considered here. Remark 5.2. If G is linear, we observe that the step (3.8) of ALG 2 is a linear problem related to the self-adjoint operator B'B. Therefore, in the finitedimensional case, assuming B injective, it will be convenient to factorize (by a Cholesky method, for example) the symmetric positive-definite matrix B'B once and for all, before starting the iterations of ALG 2. 5.4.2. On the choice of p and r If r is given, our computational experience seems to indicate that the best choice for p is p = r. The choice of r is not clear, and ALG 2 appears to be more sensitive to the choice of r than ALG 1. By the way, ALG 1 seems to be more robust on very stiff problems than ALG 2; this means that the choice of
6. Applications
183
the parameter is less critical and that the computational time for ALG 1 may become much shorter than for ALG 2. Remark 5.3. In Remark 4.2 we have seen that if JF(^) = j\q\2 and G is linear, the sequence {«"}„ related to ALG 1 converges in two iterations (at most) if we use p = r. If we use ALG 2 with the same hypotheses on F and G, then we have convergence of {«"}„ in two iterations at most only if p = r = 1 (for any choice of {p°, A1}). This fact also confirms the greater robustness of ALG 1.
6. Applications 6.1. Bingham flow in a cylindrical pipe
This is the problem considered in Chapter II, Sec. 6 and also in Sec. 1.1 of this chapter (we recall that Q is a bounded domain of U2):
Min \\ f | Vv\2 dx + g [ \Vv\dx-
[ fv dx\
(6.1)
Then (6.1) is a particular problem (P) corresponding to H = L 2 (Q) x L2(Q),
V = HJ(Q),
B = V,
(6.2)
F(q) = ^ j \q\2 dx • 2 Jn
(6.3)
G(V)=-
(6.4)
+ g j \ q \ dx,
r fv dx.
Moreover, we have F = Fo + Ft with f« . | / i ) — _
1 \n\
Jn
/TV
(6.5)
K(q) = vq,
i ^ ) = gr f |$|dx.
(6.6)
Jn
It then follows from (6.2)-(6.6) that the various assumptions required to apply Theorems 4.1 and 5.1 are satisfied. Therefore we can solve (6.1) by ALG 1 and ALG 2. Moreover, since G is linear, the result of Gabay and Mercier [1] holds (see Remark 5.1) and ALG 2 converges if 0 < pn = p < 2r. The augmented Lagrangian S£t to be used in this case is given by
JS?r(». 1, fi = I f k l 2 dx + g f \q\ dx - f fvdx+ ^ Jn
Jn
\\Vv-q\2dx. Jn
Jn
f
Jn
p.-(Vv-q)dx (6.7)
184
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
Solution of (6.1) by ALG 1. When applying ALG 1 to the solution of (6.1), it follows from (3.2)-(3.4), (6.7) that we have X° e L 2 (Q) x L2(Q), arbitrarily given;
(6.8)
then for n > 0, -rAun=f
+ W-r -rW -pnonQ,
n
p (x) = 0 ifg>\X\x)
un\r = 0,
(6.9)
+ rWu\x)\,
A"+1 = A" + pn(Vw" -/?").
(6.11)
Solution of (6.1)ftyALG 2. We have to replace (6.8) by {p°, X1} arbitrarily given in (L 2 (Q)) 2 x (L 2 (fi)) 2 ,
(6.12)
-rAu" =f+V-Xn-rV-p"'1
(6.13)
and (6.9) by onD., u"\r = 0.
Remark 6.1. In practice, (6.8)-(6.11) and (6.12), (6.13), (6.10), and (6.11) will be applied to finite element or finite difference approximations of (6.1). It then follows from (6.9), (6.13) that it is easy to use either ALG 1 (combined with the block-relaxation method of Sec. 4.3) or ALG 2 once we have at our disposal an efficient program for solving approximate Dirichlet problems for -A. Bibliographical Comments. Numerical solutions of (6.1) by ALG 1 and ALG 2 may be found in Gabay and Mercier [1] and Fortin, Glowinski, and Marrocco [1]; in Fortin [2] we can also find an iterative method of solution of (6.1) close to ALG 2 but obtained by a different approach. 6.2. Elasto-plastic torsion of a cylindrical bar This is the problem of Chapter II, Sec. 3, also considered in Sec. 1.1 of this chapter (Q is a bounded domain of U2 in the sequel):
Min I \ f |Vv\2 dx - f fv dx , V<=K [2 J n
Jn
(6.14)
J
where K = {v\v e Hj(Q), |Vv\ < 1 a.e.}; (6.14) is a particular problem (P) corresponding to V = Hj(Q),
H = L2(Q) x L2(Q),
B = V,
(6.15)
G{v)=-[fvdx,
(6.16)
F = F0 + Fu
(6.17)
6. Applications
185
where
[
\ [q\2dx^F'0{q)
= q,
(6.18)
FM = h(q)
(6.19)
with K = {q e H, \ q \ < 1 a.e.} and /^ the indicator functional of K, i.e.,
°
^!f
(6.20)
+ oo ifq$K. Here, too, it follows from (6.15)-(6.20) that the various assumptions required to apply Theorems 4.1 and 5.1 are satisfied. Therefore we can solve (6.14) by ALG 1 and ALG 2. Moreover, from the linearity of G, we have the convergence of ALG 2 if 0 < pn = p < 2r. In the present case, i f r is given by = - I \q\2 dx + Ij>(q) -
f fvdx fvdx + + f /p-(Vv -
v-q\2dx.
q) dx (6.21)
Solution o/(6.14) by ALG 1. From (3.2)-(3.4), (6.21) it follows that when applying ALG 1 to (6.14), we obtain X° arbitrarily given in (L 2 (Q)) 2 ;
(6.22)
then for n > 0, -rAu" = f + V • X" - rV • p" on Q, (6.23) «"lr = 0,
X"+1 = X" + Pn(Vu" - p").
(6.25)
Solution of (6.14) by ALG 2. We have to replace (6.22) by (6.12) and (6.23) by (6.13). Remark 6.1 still applies to (6.14), and numerical solutions of (6.14) by ALG 1 and ALG 2 may be found in Fortin, Glowinski, and Marrocco [1] and Gabay and Mercier [1].
6.3. A nonlinear Dirichlet problem In this section we follow Glowinski and Marrocco [ 5 ] ; let us consider 1 < s < +oo and Wl-S(Q)
= mQ)wllSW
where Q is a bounded domain of UN.
= {ve WU%Q.),
v\T = 0 } ,
186
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
Then we consider on Q the following nonlinear Dirichlet problem: -V-(|VM|S-2VW)=/,
(6.26) where feV' = W-Us'(O)(l/s + 1/s' = 1 => s' = s/(s - 1)). It can be proved (see, for instance, Glowinski and Marrocco [5]) that (6.26) has a unique solution which is also the solution of
Min
[- f | Vv\s dx - , u> .
(6.27)
We observe that Wl's(Q.) is not a Hilbert space if s ^ 2; therefore we cannot apply Theorems 4.1 and 5.1 to the iterative solution of (6.27). Nevertheless, once (6.27) has been approximated by a convenient finite element or finite difference method, it is possible to apply the above theorems (or Theorems 4.2 and 5.2) to the iterative solution of the approximate problem. For the sake of simplicity, we shall confine our study to the continuous problem, since it has simpler notation. We have V = Wl's{£l\
H = (Ls(Q)f,
H' = (LS'(Q))JV,
B = V,
F(q) = F0(q) = - f \q\> dx, F'(q) = q\q\'~2, s Jn G(v)= - < . / » . We observe that lim
^ =
o o ,
+
where
\q\s = ( [ \q\'dx)1"
= \\q\\vm-
We also have V p, q e H: (F'(q) - F'(pl q - p) > \q - p\l if s > 2, ( F ( g ) - F'(p), q - p ) > * UP
IF'(q) ~ F ' ( p ) \ , < P(|p\,
\ q ~ Pls2_s is "r \q\s)
+ \q\sf~2\q 1
\F'(.q)-F'(p)\,
i f l < s < 2 ,
- p \ , if s > 2, if 1 < s < 2,
where a, j5 are independent of p, q and are strictly positive.
(6.28) (6.29)
(6.30) (6.31)
6. Applications EXERCISE
187
6.1. Prove (6.28)-(6.31).
We refer to Glowinski and Marrocco [5] for a detailed analysis, including error estimates, of a finite element approximation of (6.26), (6.27) (see also Ciarlet [2]). From our numerical experiments it appears that solving (6.26), (6.27) if s is close to 1 (say 1 < s < 1.3) or large (say s > 5) is a very difficult task if one uses standard iterative methods; to our knowledge the only very efficient methods are ALG 1 and ALG 2 (or closely related algorithms; see Glowinski and Marrocco, loc. cit., for more details). The augmented Lagrangian j§?r to be used for solving (6.26), (6.27) is defined by
&Jv,q,fi)
= - f \qfdx s Jn
- < / » + T- { \ V v - q \ 2 d x + [ p • (Vi> - q) dx. 2 Jn
Jn
(6.32) Solution of (6.26), (6.27) by ALG 1. From (3.2)-(3.4), (6.32) it follows that when applying ALG 1 to (6.26), (6.27), we obtain: A°e(Z/(Q)f;
(6.33)
then for n > 0, -rAu" = / + V • A" - rV • p" in Q, (6.34) «"| r = 0, I p T Y + rp" = rVu" + X",
(6.35)
A" + 1 = X" + pn(Vun - p").
(6.36)
The nonlinear system (6.34), (6.35) can be solved by the block-relaxation method of Sec. 4.3, and we observe that if u" and X" are known (or estimated) in (6.35), the computation of p" is an easy task since \p"\ is solution of the singlevariable nonlinear equation ipT1
+r\pn\ = \rVun + Xn\,
(6.37)
which can easily be solved by various methods; once | p" \ is known, we obtain p" by solving a trivial linear equation (in (Ls(Ci))N). Solution of (6.26), (6.27) by ALG 2. We have to replace (6.33) by {p°, k1} e H x H'
(6.38)
and (6.34) by -rAun=f+
V-A" - rV
-p"-\ (6.39)
«"|r = 0.
188
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
Remark 6.1 still applies to (6.26), (6.27), and since G is linear, we can take 0 < pn = p < 2r if we are using ALG 2. For more details and comparisons with other methods, see Glowinski and Marrocco [5], [7] and Fortin, Glowinski, and Marrocco [1]. Remark 6.2. ALG 1 and ALG 2 have also been successfully applied to the iterative solution of magneto-static problems (see Glowinski and Marrocco [6]). They have also been applied by Fortin, Glowinski, and Marrocco [1] to the solution of the subsonic flow problem described in Chapter IV, Sec. 3; in the latter case, using ALG 1 and ALG 2, we obtain easy variants of (6.33)(6.36) and (6.38), (6.39), (6.35), and (6.36).
6.4. Application to the solution of mildly nonlinear systems Let A be a N x N symmetric positive-definite matrix, let D be a diagonal positive-semidefinite matrix, and let f e UN. Let >: R -> R be a C° and nondecreasing function (we can always suppose that 0(0) = 0). Using the same notation as in Chapter IV, Sec. 2.6, we associate to v = {vu v2, •. •, vN} e UN the vector
Vi=l,...,N.
(6.40)
Then we consider the nonlinear system Au + D0(u) = f.
(6.41)
In Chapter IV, Sec. 2.6, various methods for solving (6.41) have been given, but in this section we would like to show that (6.41) can also be solved by ALG 1 and ALG 2 once a convenient augmented Lagrangian has been introduced. Remark 6.3. The methods to be described later are easily generalized to the case where A is not symmetric but still positive definite. Let us define
Jo
Since cj) is C° and nondecreasing, we find that $ is C 1 and convex. It then follows from the symmetry of A that solving (6.41) is equivalent to solving the minimization problem J(u) < J(v),
V v e UN, u e UN.
(6.42)
In (6.42) we have
J(y) = \ (Av, v) + £ dMvd - (f, v),
(6.43)
6. Applications
189
where (-, •) denotes the usual inner product of UN and || • || denotes the corresponding norm and where
D =|
d,
From the above properties of A, D, and $, it follows from, e.g., Cea [1], [2], that (6.41), (6.42) has a unique solution. Remark 6.4. In fact (6.41) has a unique solution if A is positive definite and possibly nonsymmetric, the assumptions on
(6.44)
G(v)= £ d, <*(»,)-(f,v),
(6.45)
F(q) = F 0 (q) = ^Aq, q) => F 0 (q) = Aq.
(6.46)
From these properties we can solve (6.41), (6.42) by using ALG 1 and ALG 2 (we observe that, unlike in the above examples, G is nonlinear). Remark 6.5. Instead of using G and F defined by (6.44), (6.45), we can use
GOO = £ dMvd, F(q) = i(Aq, q) - (f, q). The augmented Lagrangian to be associated with (6.44)-(6.46) is ^,(v, q, »i) = \ (Aq, q) + £ d,
(6.48)
190
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
then for n > 0, ru" + D
(6.49)
(rl + A)p" = ru" + X",
(6.50)
The nonlinear system (6.49), (6.50) can be solved by the block-relaxation method of Sec. 4.3, and we observe that if p" and X" are known (or estimated) in (6.49), the computation of u" is easy since it is reduced to the solution of N independent single-variable nonlinear equations of the following type: r£ + #(
(6.52)
Since r > 0 and 4> is C° and nondecreasing, (6.52) has a unique solution which can be computed by various standard methods (see, e.g., Householder [1] and Brent [1]). Similarly, if u" and X" are known in (6.50), we obtain p" by solving a linear system whose matrix is rl + A. Since r is independent of n, it is very convenient to prefactorize rl + A (by the Cholesky or Gauss method). Solution of (6.41), (6.42) by ALG 2. We have to replace (6.48) by {p°, X1} eUN xUN
(6.53)
and (6.49) by ru" + D
(6.54)
From Theorem 5.2 it follows that we have convergence of (6.53), (6.54), (6.50), and (6.51) if 0 < Pn = p < (1 + y/S)/2r. Remark 6.6. Suppose that pn = p = r in ALG 2; we then have ru" + D0(u") = f + rp"" 1 - Xn, rp" + Ap" = ru" + X",
(6.55)
Xn+1 = X" + r ( u " - p " ) . From (6.55) it follows that X"+1=Ap".
(6.56)
Then from (6.55), (6.56) we obtain ru" + D>(u") + Ap"- J = f + rp"- \ rp" + Ap" + D>(u") = f + rp"" 1 .
(6.57) (6.58)
Therefore, if pn = p = r, ALG 2 reduces (with different notation) to the alternating-direction method described in Chapter IV, Sec. 2.6.6. (for more details on the relation existing between alternating-direction methods and
6. Applications
191
augmented Lagrangian methods, we refer to G.L.T. [3, Appendix 2], Gabay [1], Bourgat, Dumay, and Glowinski [1], and Bourgat, Glowinski, and Le Tallec [1]). Remark 6.7. From the numerical experiment performed in Chan and Glowinski [1], ALG 1, combined with the block-relaxation method of Sec. 4.3, is much more robust than ALG 2; this is the case if, for instance, we solve a finite element (or finite difference) approximation of the mildly nonlinear elliptic problem -Aw + M|M| S " 2 = /
onQ,
« | r = 0,
(6.59)
with 1 < s < 2. In Chan and Glowinski, loc. cit., we can find various numerical results and also comparisons with other methods (see also Chan, Fortin, and Glowinski [1]).
6.5. Solution of elliptic variational inequalities on intersections of convex sets 6.5.1. Formulation of the problem Let V be a real Hilbert space and a:V x F->[Rbea bilinear form, continuous, symmetric, and F-elliptic. Let K be a closed convex nonempty subset of V such that N
K= f]Kh
(6.60)
i=l
where, V i = l,...,N,Ktisa. EVI problem
closed convex subset of V. We then consider the
a(u, v - u) > L(v - u),
VveK,
us K,
(6.61)
where L: V —>• U is linear and continuous. Since a(-, •) is symmetric, we know from Chapter I that the unique solution of (6.61) is also the solution of J(u) < J(v),
VceX,
ueK,
(6.62)
where J(v) = \a(v,v) - L(v).
(6.63)
6.5.2. Decomposition of (6.61), (6.62) Let us define (with q = {qu...,
qN})
W = {{v, q} e V x VN, v - qt = 0,
V i = 1 , . . . , JV}
(6.64)
192
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
and Jf = {{v,q}eW,qieKi,
V i = 1,..., N}.
(6.65)
It is clear that (6.62) is equivalent to Min j(v, q),
(6.66)
where
Kv, «) = ^
I «(«*. 9d - UP)-
(6-67)
Remark 6.8. We have to observe that many other decompositions are possible; for instance, W = {{v, q}eV
x VN,v - ql=0,qi+1-
qt = 0,
V i = 1 , . . . , N - 1}
with j and X again defined by (6.67), (6.65). We can also use W = {{v,q}eV
x VN~\ v - qt = 0,
V i = 1,... ,N - 1}
with JtT = {{v,q} eW,ve
Kuqie
Ki+U
Vi = 1.....N - 1}
and 1 1 N~l J(P,
Bv = {v,..., D},
G(v) = -L(i>),
f o(«)•= ^
t
a
^ id,
^i(«) = t IM
(6.68) (6.69)
(6.70) (6.71)
i=l
with IK.: indicator function of Kt. It is easily shown that from the properties of B, G, F we can apply ALG 1 and ALG 2 to solve (6.62), via (6.66), provided that the augmented Lagrangian r N 1N £Cr(v, q, n) = F(q) + G(v) + — J] a(v - qt, v - qt) + — J] Qi{, v - qt) (6.72)
6. Applications
193
has a saddle point over V x VN x VN. Such a saddle point exists if H is finite dimensional, since the constraints v — qi = 0 are linear. 6.5.3. Solution of (6.62) by ALG 1. From (3.2)-(3.4), (6.72) it follows that when applying ALG 1 to (6.62), we obtain: X° then for n >
given;
(6.73)
o, XI, v ) +L(v),
- 1
=r ra(u", v) =
MveV,
u"eV, (6.74)
(1 + r)a(p1,
- pi) > ra(u",
-Pi) + (XI, q<-Pi),
V qt e Kh
pi e Kt (6.75),
/ori=l,2,...,JV; ^ + 1 = ^ + P«(«"-P?)
(6.76)i
/or i = 1 , . . . , AT. The system (6.74), (6.75) is, for X" given, a system of coupled EVI's; a very convenient method for solving it is the block-over-relaxation method with projection described in Chapter V, Sec. 5 and also in Cea and Glowinski [1] and Cea [2]. This method will reduce the solution of (6.62) to a sequence of EVI's on Kh i = 1 , . . . , N. 6.5.4. Solution of (6.62) by ALG 2 From (3.7)-(3.10), (6.72) it follows that to solve (6.62) by ALG 2, we have to use the variant of (6.73)-(6.76) obtained by replacing (6.73), (6.74) with: {p°,X1}eVN
ra(u\ v) = ra(~ |
P1~
x VN given;
\ v\ - l~ | XI, v\ + L(v),
(6.77)
V v e V, u" e V. (6.78)
Remark 6.9. The two algorithms above are well suited for use in multiprocessor computers, since many operations may be done in parallel; this is particularly clear for algorithm (6.77), (6.78), (6.75), (6.76). Remark 6.10. Using different augmented Lagrangians, other than <£r denned by (6.72), we can solve (6.62) by algorithms better suited to sequential computing than to parallel computing. We leave to the reader, as exercises, the task of describing such algorithms.
194
VI Decomposition-Coordination Methods by Augmented Lagrangian: Applications
Remark 6.11. The two algorithms described above can be extended to EVI's where a(-, •) is not symmetric. Moreover, they have the advantage of reducing the solution of (6.62) to the solution of a sequence of simpler EVI's of the same type, to be solved over Kt, i = 1,...,N, instead of K.
7. General Comments As mentioned several times before, the methods described in this chapter may be extended to variational problems which are not equivalent to optimization problems. These methods have been applied by Begis and Glowinski [1] to the solution of fourth-order nonlinear problems in fluid mechanics (see also Begis [2] and G.L.T. [3, Appendix 6]). From a historical point of view, the use of augmented Lagrangians for solving—via ALG 1 and ALG 2—nonlinear variational problems of type (P) (see Sec. 1.1) seems tobedueto Glowinski and Marrocco [3], [4], [5]. For more details and other applications, see Gabay and Mercier [1], Fortin and Glowinski [1], [2], Glowinski and Marrocco, loc. cit., and also Bourgat, Dumay, and Glowinski [1], Glowinski and Le Tallec [1], [2], Le Tallec [1], Bourgat, Glowinski, and Le Tallec [ 1 ], and Glowinski, Le Tallec, and Ruas [ 1 ], where ALG 1 and ALG 2 have been successfully used for solving nonlinear nonconvex variational problems occurring in finite elasticity (particularly in inextensible and/or incompressible finite elasticity). With regard to Sec. 3.2, D. Gabay [1] has recently introduced the following variant of ALG 2: {p°, Xl}eH n
n
x H given; n+1 2
then, {p ~\ X"} known, we define {u , X n
(7.1)
n+1
' , p", X
n
} by
n
G(v) - G(u ) + (X , B(v - «")) + r(Bu" - p ~ \ B(v - if)) > 0, V v e V, uneV, 1/2
y +
F(q) - F(p") - (A"
+1/2
, q-
Xn +
=
n
p(Bun
_
pn-
n
1^
(73) n
p )+ r(p" - Bu , q - p ) > 0, \fqeH,
+J
A"
(7.2)
= X
n+1/2
+ p(Bu" - p"),
p"eH,
(7.4) (7.5)
with p > 0 in (7.3), (7.5). For additional details and convergence properties, see Gabay, loc. cit. (and also Gabay [2]). To conclude this chapter, we have to mention that, using some results due to Opial [1], we have, in fact, in Theorems 4.1 and 5.1 (resp., 4.2 and 5.2) the weak convergence (resp., the convergence) of the whole sequence {X"}n to a A* such that {u, p, X*} is a saddle point of i f (and &r) over V x H x H. We refer to Glowinski, Lions, and Tremolieres [3, Appendix 2] for a proof of the above results in a more general context (see also Gabay [2]).
CHAPTER VII
Least-Squares Solution of Nonlinear Problems: Application to Nonlinear Problems in Fluid Dynamics
1. Introduction: Synopsis In this chapter we would like to discuss the solution of some nonlinear problems in fluid dynamics by a combination of least-squares, conjugate gradient, and finite element methods. In view of introducing the reader to this technical subject, we consider in Sec. 2 the solution of systems of nonlinear equations in UN by least-squares methods; then, in Sec. 3, the solution of a nonlinear Dirichlet model problem; also in Sec. 3 we make some comments about the use of pseudo-arc-length-continuation methods for solving nonlinear problems. In Sec. 4 we discuss the application of the above methods to the solution of the nonlinear equation modelling potential transonic flows of inviscid compressible fluids; finally in Sec. 5 we discuss the solution of the Navier-Stokes equations, for incompressible viscous Newtonian fluids, by similar techniques. This chapter is closely related to Bristeau, Glowinski, Periaux, Perrier, and Pironneau [1] and Bristeau, Glowinski, Periaux, Perrier, Pironneau, and Poirier [1]; other references will be given in the sequel.
2. Least-Squares Solution of Finite-Dimensional Systems of Equations 2.1. Generalities Replacing the solution of finite-dimensional systems of equations by the solution of minimization problems is a very old idea, and many papers dealing with this approach can be found in the literature. Since referring to all those papers would be an almost impossible task, we shall mention just some of them, referring to the bibliographies therein for more references. The methods most widely used have been the least-squares methods in which the solution of F(x) = 0, where F: UN ->• UN with F = {/,,..., fN}, is replaced by:
(2.1)
196
VII Least-Squares Solution of Nonlinear Problems
Find xeUN such that
VyeR",
(2.2)
where in (2.2), || • || denotes some Euclidean norm. If N is not too large, a natural choice for || • || is (if y = {yu ..., yN}) /N
\l/2
nyi = ( w )
•
(23)
Suppose, for example, that F(x) = Ax - b,
(2.4)
where A is an N x N matrix and b e UN. If || • || is denned by (2.3), then the corresponding problem (2.2) is equivalent to the well-known normal equation ArAx = A'b,
(2.5)
where A' is the transpose matrix of A. This simple example shows the main advantage of the method, which is to replace the original problem Ax = b,
(2.6)
whose matrix is possibly nonsymmetric and indefinite, by the problem (2.5) whose matrix is symmetric and positive semidefinite (or equivalently, by the minimization of a quadratic convex functional). This convexification property (which can only be local in nonlinear problems) is fundamental since it will insure the good behavior (locally, at least) of most minimization methods used to solve the least-squares problem (2.2) (once a proper || • || has been chosen; see below). Also, from (2.5) it is clear that a main drawback of the method is the possible deterioration of the conditioning which, for example, may make the solution of (2.2) sensitive to roundoff errors. Actually in many problems this drawback can be easily overcome by the use of a more sophisticated Euclidean norm than (2.3). Indeed, if || • || is defined by llyll = (Sy, y)^ 2
(2.7)
(where S is an N x N positive-definite symmetric matrix and (x, y)RW = Xf= i xiyi)an^ if F is still defined by (2.4), then (2.5) is replaced by A'SAx = A'Sb.
(2.8)
With a proper choice of S we can dramatically improve the conditioning of the matrix in the normal equation (2.8) and make its solution much easier. This matrix S can be viewed as a scaling (or preconditioning) matrix. This idea of preconditioning stiff problems will be systematically used in the sequel. The standard reference for linear least-squares problems is Lawson and Hanson [1]; concerning nonlinear least-squares problems of finite dimension
2 Least-Squares Solution of Finite-Dimensional Systems of Equations
197
and their solution, we shall mention, among many others, Levenberg [1], Marquardt [1], Powell [3], [4], Fletcher [1], Golub and Pereyra [1], Golub and Plemmons [1], Osborne and Watson [1], and More [1] (see also the references therein).
2.2. Conjugate gradient solution of the least-squares problem (2.2) Conjugate gradient methods have been considered in Chapter IV, Sec. 2.6.7; actually they can also be used for solving the least-squares problem (2.2). We suppose that in (2.2) the Euclidean norm ]| • || is defined by (2.7), with S replaced by S~1, and we use the notation /
(x, y) = (x, y)R»
JV
I = X,xf
Let us define J : UN -^ R by (2.9) we clearly have equivalence between (2.2) and the following: Find xeUN such that J(x) < J(y),
VyeR"
(2.10)
In the following we denote by F' and J' the differentials of F and J, respectively, we can identify F' with the (Jacobian) matrix (3/;/<3XJ)I<;,J<;V> a n d we have (J'(y), z) = ( S - ^ ( y ) , F'(y)z),
Vy.zeR"
(2.11)
which implies (2.12) To solve (2.2) (via (2.10)), we can use the following conjugate gradient algorithm in which S is used as a scaling (or preconditioning) matrix (most of the notation is the same as in Chapter IV, Sec. 2.6.7). First algorithm (Fletcher-Reeves) x°eUN given;
(2.13)
g^S-W),
(2.14)
w° = g°.
(2.15)
Then assuming that x" and w" are known, we compute x" + 1 by x" + 1 = x" - pnW,
(2.16)
where pn is the solution of the one-dimensional minimization problem J(x" - pBw") < J(x" - pw"),
V p e R, pneU.
(2.17)
198
VII Least-Squares Solution of Nonlinear Problems
Then g«+i
and compute w
B+1
= S" 1 J'(x" +1 )
(2.18)
by W+1 = g"+1 + /lnw",
(2.19)
where (Sg" +1 g" +1 ) (220) "(Sg«,g") • Second algorithm (Polak-Ribiere). This method is like the first algorithm except that (2.20) is replaced by A
g) (221) • Remarks 2.17, 2.19, and 2.20 of Chapter IV, Sec. 2.6.7 still hold for algorithms (2.13)-(2.20) and (2.13)-(219), (2.21). As a stopping test for the above conjugate gradient algorithms, we may use, for example,either J(x") < £or||g"|| Si e (where eis a "small"positive number), but other tests are possible.
3. Least-Squares Solution of a Nonlinear Dirichlet Model Problem In order to introduce the methods that we shall apply in Sees. 4 and 5 to the solution of fluid dynamics problems, we shall consider the solution of a simple nonlinear Dirichlet problem by least-squares and conjugate gradient methods after briefly describing (in Sec. 3.2) the solution of the model problem introduced in Sec. 3.1 by some more standard interative methods; in Sec. 3.5 we shall briefly discuss the use of pseudo-arc-length-continnation methods for solving nonlinear problems via least-squares and conjugate gradient algorithms. 3.1. Formulation of the model problem Let Q c UN be a bounded domain with a smooth boundary F = dQ; let T be a nonlinear operator from V = Hj(Q) to V* = H " 1 ^ ) {H~l(£l): topological dual space of //J(Q)). We consider the nonlinear Dirichlet problem: Find u e HQ(Q) such that ~Au - T(u) = OinQ,
(3.1)
and we observe that u e HQ(Q) implies u = 0 on T. Here we shall not discuss the existence and uniqueness properties of the solutions of (3.1), since we do not want to be very specific about the operator T.
3 Least-Squares Solution of a Nonlinear Dirichlet Model Problem
199
3.2. Review of some standard iterative methods for solving the model problem 3.2.1. Gradient methods The simplest algorithm that we can imagine for solving (3.1) is as follows: u° given; +1
then for n > 0, define u"
(3.2)
from u" by -Au"+l
= T(un) in Q, +1
»"
(i
= 0onr.
'
Algorithm (3.2), (3.3) has been extensively used (see, e.g., Norrie and De Vries [1] and Periaux [1]) for the numerical simulation of subsonic potential flows for compressible inviscid fluids like those considered in Chapter IV, Sec. 3. Unfortunately algorithm (3.2), (3.3) usually blows up in the case of transonic flows. Actually (3.2), (3.3) is a particular case of the following algorithm: u° given; +i
then for n > 0, define u"
(3.4)
from u" by _A« B + 1 / 2 = T(u") in fi, M"
+1/2
(3
= 0onr,
u n + 1 = u" + p ( u n + 1 ' 2 - u"),
p > 0 .
-^
(3.6)
If p = 1 in (3.4)-(3.6), we recover (3.2), (3.3). Since (3.5), (3.6) are equivalent to = „" - p ( - A ) - 1 ( - A « " - T(u")), (3.7) with ( — A)" corresponding to Dirichlet boundary conditions, algorithm (3.4)-(3.6) is very close to a gradient method (and is rigorously a gradient algorithm if T is the derivative of some functional). Let us define A: Hj(fi) -> H~ \Q) by M-+i
1
A(v)=
-Av-
T(v).
(3.8)
We can easily prove the following: Proposition 3.1. Suppose that the following properties hold (i) A is Lipschitz continuous on the bounded sets ofHl(Q); (ii) A is strongly elliptic, i.e., there exists a > 0 such that a\\v2-vl\\2Hh(n),
V vu v2 eH^Q)
(3.9)
(where < •, •) denotes the duality between H~ 1(Q) and HQ(Q)). Then problem (3.1) has a unique solution; moreover, for every u° e Hl(Q), there exists pM > 0 (depending upon u° in general) such that 0 < p < pM
(3-10)
implies the strong convergence o/(3.4)-(3.6) to the solution u o/(3.1). PROOF. See, e.g., Brezis and Sibony [2].
•
200
VII Least-Squares Solution of Nonlinear Problems
3.2.2. Newton's methods Assuming that T is differentiable, one may try to solve (3.1) by a Newton's method. For this case, using a prime to denote differentiation, we obtain: u° given; +i
(3.11)
n
then for n > 0, define ii" from u by - AM"+ 1 - T'(u") • un+1 = T(un) - T'(nn) • u" in Q, un+1 =0 on T. (3.12) Algorithm (3.11), (3.12) is the particular case corresponding to p = 1 of the following: u° given;
(3.13)
+1
then for n > 0, define n" from u" by - Aun+1/2 -
T"(M")
• "" + 1 / 2 = T(un) - T'(u") • u" in Q, un+m = 0 on T, (3.14)
n+1
u
n
n+112
=u + p(u
-u"),
p>0.
(3.15)
The various comments made in Chapter IV, Sec. 2.6.4 about Newton's methods still hold for algorithms (3.11), (3.12) and (3.13)-(3.15). 3.2.3. Time-dependent approach A well-known technique is the following: one associates with (3.1) the timedependent problem ~ -
AM - 7X10 = 0 infi,
(3.16)
ot « = 0onr,
(3.17)
u(x, 0) = uo(x) (initial condition). (3.18) Since lim,^ + 00 u(t) are usually solutions of (3.1), a natural method for solving (3.1) is the following: (i) Use a space approximation to replace (3.16)—(3.18) by a system of ordinary differential equations, (ii) Use an efficient method for numerically integrating systems of ordinary differential equations, (iii) Then integrate from 0 to + oo (in practice, to a large value of t). In the case of a stiff problem, it may be necessary to use an implicit method to integrate the initial-value problem (3.16)-(3.18). Therefore each time step will require the solution of a problem like (3.1). If one uses the ordinary backward implicit scheme (see Chapter III, Sec. 3) one obtains u° = u0,
(3.19)
3 Least-Squares Solution of a Nonlinear Dirichlet Model Problem
201
and for n > 0, w " + 1 - it"
k
- AM"+ 1 - T(un+!) = 0 in Q,
un+1 = 0 on Y
(3.20)
(where k denotes the time-step size). At each step one has to solve
k
+! - Au" +1 - T(u" ) = \ in Q, un+1 = 0 on F, v ' k
(3.21)
which is very close to (3.1) (but usually better conditioned); actually, in practice, instead of (3.21), we solve a finite-dimensional system obtained from (3.16) (3.18) by a space discretization. 3.2.4. Alternating-direction methods These methods have been considered in Chapter IV, Sec. 2.6.6; actually they are also closely related to the time-dependent approach as can be seen in, e.g., Lions and Mercier [1] to which we refer for further results and comments. Two possible algorithms are the following: First Algorithm. This is a nonlinear variant of the Peaceman-Rachford algorithm (see Peaceman and Rachford [1], Varga [1], Kellog [1], Lions and Mercier [1], and Gabay [1]) defined by: u° given;
(3.22) +1/2
then for n > 0, u" being given, we compute u" rnun+1/2
- T(un+1/2)
+1
, u"
from u" by
= rnu" + AM",
rnun+1 - A«" +1 = rnun+m
+ T(u"+'12).
(3.23) (3.24)
Second Algorithm. This is a nonlinear variant of the Douglas-Rachford algorithm (see Douglas and Rachford [1], Lieutaud [1], Varga, Lions and Mercier, and Gabay, loc. cit.) denned by (3.22) and rnun+li2 n+1
rnii
~
T(M"+1/2)
- A«"
+1
= rnu" + Aun,
= rnu" + T(«"
+ 1/2
).
(3.25) (3.26)
In both algorithms {rn}n > 0 is a sequence of positive parameters (usually a cyclic sequence). In the nonlinear case, the determination of optimal sequences {/«}«> o i s a difficult problem. 1 We also have to observe that if in algorithm
1 See, however, Doss and Miller [1] in which alternating-direction methods more sophisticated than (3.22H3.24) and (3.22), (3.25), (3.26) are also discussed and tested.
202
VII Least-Squares Solution of Nonlinear Problems
(3.22)-(3.24) operators T and A play the same role, this is no longer true in (3.22), (3.25), (3.26), and it is usually safer to have A as an "acting" operator in the second step (if we suppose that —A is "more" elliptic than — T).
3.3. Least squares formulations of the model problem (3.1) 3.3.1. Generalities We shall consider least-squares formulations of the model problem (3.1). An obvious least-squares formulation consists of the statement that the required function u minimizes the left-hand side of (3.1) in a L2(Q)-least-squares sense. That is, Min f |Ai> + T(v)\2dx,
(3.27)
•>£!
veV
where V is a space of feasible functions. Let us introduce £ by - A £ = T(t>)inn,
(328)
£ = 0 on F. Then (3.27) is equivalent to Min f |A(t>- O\2dx, veV
(3.29)
Jn
where £ is a (nonlinear) function of v, through (3.28). From Lions [4] and Cea [1], [2], for example, it is clear that (3.28), (3.29) Has the structure of an optimal control problem where (i) (ii) (iii) (iv)
v is the control vector, £ is the state vector, (3.28) is the state equation, and the functional occurring in (3.29) is the cost function.
Another least-squares optimal control formulation is
Min f \ v - £\2 dx, veV
(3.30)
Jn
where £ again satisfies (3.28). This formulation has been used by Cea and Geymonat [1] to solve nonlinear partial differential problems (including the steady Navier-Stokes equations). Actually the two above least-squares formulations may lead to a slow convergence, since the norm occurring in the cost functions is not appropriate for the state equation. An alternate choice, very well suited to nonlinear second-order Dirichlet problems, will be discussed in the next section.
3 Least-Squares Solution of a Nonlinear Dirichlet Model Problem
203
3.3.2. A H~i-least-squares formulation of (3.1) Let us recall some properties of H~ ^Q), the topological dual space of If L 2 (Q) has been identified with its dual space, then
HQ(Q).
moreover A ( = V2) is an isomorphism from //o(Q) onto H~1(Q). In the sequel the duality pairing < •, • > between H~X(H) and //J(Q) is chosen in such a way that V/eL2(Q),
< / , » > = [fvdx, Jo
VUGHJ(Q).
(3.31)
The topology of H~^Q) is defined by || • ||*, where, V / e i f " \ Q ) , SU
=
P
liTij l\vl\^
(3-32>
•
A convenient2 least-squares formulation for solving the model problem (3.1) seems to be
Min | | A » + TO;) II*.
(3.33)
It is clear that if (3.1) has a solution, then this solution will be a solution of (3.33) for which the cost function will vanish. Let us introduce £ e Hj(Q) by A£ = Av + T(v) in £1, t = 0 on F so that (3.33) reduces to Min ||A£|L,
(3.35)
where £, is a function of v through (3.34). Actually it can be proved that if || • || „. is defined by (3.32) with < •, • > obeying (3.31), then VveHKtt).
(3.36)
From (3.36) it then follows that (3.35) may also be formulated by Min
2
\\Vt;\2dx
(3.37)
Convenient because the space H\,{£i) in (3.33) is also the space in which we want to solve (3.1) (from the properties of A and T).
204
VII Least-Squares Solution of Nonlinear Problems
where ^ is a function of v through (3.34); (3.37) also has the structure of an optimal control problem. Remark 3.1. Nonlinear boundary-value problems have been treated by Lozi [1] using a formulation closely related to (3.34), (3.37).
3.4. Conjugate gradient solution of the least squares problem (3.34), (3.37) Let us define J : i?£(fi) ->• U. by dx,
(3.38)
where ^ is a function of v in accordance with (3.34); then (3.37) may also be written as follows: Find u e HQ(Q) such that J(u) < J(v),
V c e f f J(Q).
(3.39)
To solve (3.39), we shall use a conjugate gradient algorithm. Among the possible conjugate gradient algorithms, we have selected the Polak-Ribiere version (see Polak [1]), since this algorithm produced the best performances in the various experiments that we performed (the good performances of the Polak-Ribiere algorithm are discussed in Powell [2]). Let us denote the differential of J(-) by J'( •); then the Polak-Ribiere version of the conjugate gradient method applied to the solution of (3.39) is Step 0: Initialization u° e HJ(Q) given;
(3.40)
then compute g° e Ho(Q)/rom (
)
g° = 0 on T, and set z° = g°.
(3.42) +1
+1
+i
Then for n > 0, assuming u", g", z" known, compute u" , g" , z"
by:
Step 1: Descent un+1 =un-
Xnz\
(3.43)
where Xn is the solution of the one-dimensional minimization problem Xn e U,
J(un - V ) < J{ii" - Az"),
V X e U.
(3.44)
3 Least-Squares Solution of a Nonlinear Dirichlet Model Problem
205
Step 2: Construction of the New Descent Direction. Define g"+1 eHj(Q) by -Agn+1
= J'(un+1) in Q,
0" + 1 = O o n r ;
(3.45)
then
z*+1=g»+1+yHz»,
(3.47)
n = n + 1 go to (3.43).
(3.48)
The two nontrivial steps of algorithm (3.40)-(3.48) are: (i) The solution of the single-variable minimization problem (3.44); the corresponding line search can be achieved by dichotomy or Fibonacci methods (see, for example, Polak [1], Wilde and Beightler [1], and Brent [1]). We have to observe that each evaluation of J(v) for a given argument v requires the solution of the linear Dirichlet problem (3.34) to obtain the corresponding £. (ii) The calculation of gn+1 from un+1 which requires the solution of two linear Dirichlet problems (namely (3.34) with v = M" +1 and (3.45)). Calculation ofJ'(u") and g". Due to the importance of step (ii), let us describe the calculation of J'(u") and g" in detail. Let weflJ(Q); then J'{v) may be denned by (3.49) (-•0
(*0
from (3.34), (3.38), (3.49) we obtain <J'(»),w>= [v^-Vndx,
(3.50)
Ja
where t] e Hl(Q) is the solution of AM = Aw + T'(v) • w in Q,
(3.51) n = 0 on F; (3.51) has the following variational formulation: (Vn-Vzdx
= f Vw-Vzdx -
Vze#J(Q),
f/eff&fi). (3.52)
206
VII Least-Squares Solution of Nonlinear Problems
Taking z = £ in (3.52), from (3.50) we obtain <J'(v), w> = f V£ • Vw dx - < 7 » • w, O ,
V », w E tf J(fl).
(3.53)
Therefore J'(v) (eH~l(Q.)) may be identified with the linear functional on Hj(Q) defined by -
(3.54)
From (3.45), (3.53), (3.54) it then follows that g" is the solution of the following linear variational problem: Find g" e H£(Q) such that, V w e i/£(fi), f Vg" • Vw dx = f V
(3.55)
where for w belonging to a basis of the finite dimensional subspace of HQ(Q.) corresponding to the Galerkin or finite element approximation under consideration. 3.1. Extend the above least-squares-conjugate-gradient methodology to the solution in HQ(Q) X H J(Q) of the nonlinear boundary-value system
EXERCISE
-
AMJ
du, dut , . n + ut — - + « 2 ^— = / i m £2>
Ai +u +u LmQ
(156)
- '> ^ >tr ' Uj = u2 = 0 on F,
where Q is a bounded domain of K2 and f1,f2eH~
'(Q).
3.5. A nonlinear least-squares approach to arc-length-continuation methods 3.5.1. Generalities: Synopsis We would like to show that the above least-squares methodology can be (slightly) modified in order to solve nonlinear problems by arc-lengthcontinuation methods, directly inspired by H. B. Keller [1], [2] (where the
3 Least-Squares Solution of a Nonlinear Dirichlet Model Problem
207
basic iterative methods are Newton's and quasi-Newton's, instead of conjugate gradient). As a test problem we have chosen a variant of the nonlinear Dirichlet problem (3.1); let us consider the following family of nonlinear Dirichlet problems, 3 parametrized by X e U, -A« = XT(u) in Q, u = OonT;
(3.57)
(3.1) corresponds to X = 1. 3.5.2. Solution of(3.57) via arc-length-continuation methods Following H. B. Keller [1], [2] [for which we refer for justification (see also Percell [1])], we associate to (3.57) a " continuation " equation; we have chosen 4 |VM| 2 dx + X2 = 1,
(3.58)
where u = du/ds, X = dXjds, and where the curvilinear abscissa s is defined by Ss = XSX + f VM • VSu dx,
(3.59)
(5s)2 = (5Xf + f VSu • Wdu dx.
(3.60)
or equivalently by
In fact we are considering a path in Hj(Q) x IR whose arc length is defined by (3.58)-(3.60). Then, in order to solve (3.57), we consider the family (parametrized by s) of nonlinear systems (3.57), (3.58). In practice we shall approximate (3.57), (3.58) by the following discrete family of nonlinear systems, where As is an arc-length step, positive or negative (possibly varying with n) and u" ~ u(n As): Take u° = 0, X° = 0 and suppose that 1(0), «(0) are given
(3.61)
(initialization (3.61) is justified by the fact that u = 0 is the unique solution of (3.57) if X = 0); then for n > 0, assuming M"" 1 , X"~l, u", X" known, we obtain {un+1, Xn+1}eHl(Q.) x Uby -Aun+1
=
Xn+1T(un+1)inQ,
un+1 = 0onY,
3 4
T o be solved in j() O t h e r choices are possible
(3.62)
208
VII Least-Squares Solution of Nonlinear Problems
and f Viu1 - u°) • Vu(0) dx + (X1 - A°)l(0) = As ifn = 0,
= 0 or 1, ifn>\.
(3.63)
(3.64)
Obtaining ti(0) and 1(0) is an easy task since we have (from (3.57))
«(0) = 0 on T,
(3.65)
and therefore
X2(0)(l + j \Vu\2 dx\ = I,
(3.66)
where M e //J(Q) is the solution of -Aw = T(0)infi, M=
0 on T
(3.67)
(then we clearly have a(0) = X(0)u). Relations (3.61)-(3.64) look like clearly a discretization scheme for solving the Cauchy problem for first-order ordinary differential equations; from this analogy we can derive many other discretization schemes for the approximation of (3.57), (3.58) (Runge-Kutta, multisteps, etc...) and also methods for the automatic adjustment of As. 3.5.3. Nonlinear least squares and conjugate gradient solution o/(3.62)-(3.64). Without going into details (for which we refer to Reinhart [1] and Glowinski, Keller, and Reinhart [1]) we can solve (3.62)-(3.64) by a variant of (3.40)-(3.48) defined on the Hilbert space Hl(Q) x IR equipped with the metric and innerproduct corresponding to (3.68) Jn it is clear that many other norms than (3.68) are possible, however in all cases the scaling of a conjugate gradient algorithm using a discrete variant of5 -A
0\ I (or similar operators)
will require an efficient solver and the conclusions of Sec. 3.4 still hold. 1
In (3.69), A corresponds to the homogeneous Dirichlet boundary conditions.
(3.69)
3 Least-Squares Solution of a Nonlinear Dirichlet Model Problem
209
Remark 3.4. To initialize the conjugate gradient algorithm solving (3.62)(3.64) we have used {21" - A""1, 2M" - M"" 1 } as initial guess to compute {Xn+1, un+1}; with such a choice we obtain a much faster convergence than by taking {A", «"} as initial guess. 3.5.4. Application 3.5.4.1. Formulation of the problem. We shall apply the methods described in Sees. 3.5.2 and 3.5.3 to the solution of the following classical problem: - A M = Ae"infi, (170) n r u = 0 on F, where Q is a bounded domain of UN. We have to observe that unless N = 1, the mapping T denned by
is not continuous from HQ(O) to V* ( = H~1(Q)). We consider only the case where A > 0, since if X < 0, the operator v -* — Av — Xe" is monotone and therefore the methods of Chapter IV, Sec. 2 can be applied (take (j)(i) = — X(e' — 1 ) , / = A), showing the existence of a unique solution of (3.70) (which is u = 0 if A = 0). With A > 0, problem (3.70) has been considered by many authors (Henri Poincare—with Q = IRN—among them). With regard to recent publications, let us mention, among others, Crandall and Rabinowitz [1], [2], Amann [1], Mignot and Puel [1], and Mignot, Murat, Puel [1]. In particular, in Mignot, Murat, and Puel [1] we may find an interesting discussion showing the relationships between (3.70) and combustion phenomena. From a numerical point of view, problem (3.70) has been investigated by, among others, Kikuchi [1] and Reinhart [1] to which we refer for more details and further references (see also Simpson [1], Moore and Spence [1], Glowinski, Keller, and Reinhart [1], and Chan and Keller [1]). 3.5.4.2. Numerical implementation of the methods of Sees. 3.5.2. and 3.5.3. We have chosen to solve the particular case of (3.70) where Q = ]0, 1[ x ]0, 1[. The practical application of the methods of Sees. 3.5.2 and 3.5.3 requires the reduction of (3.70) to a finite-dimensional problem; to do this we have used the finite element method described in Chapter IV, Sec. 2.5, taking for ^h the triangulation consisting of 512 triangles indicated in Fig. 3.1. The unknowns are the values taken by the approximate solution uh at the interior nodes of 2Th; we have 225 such nodes. Algorithm (3.61)-(3.64) has been applied with As = 0.1, and) = 0 in (3.64); we observe that T(0) = 1 in (3.67); algorithm (3.61)-(3.64) ran "nicely," since an accurate least-squares solution of the nonlinear system (3.62), (3.64) required basically no more than 3 or 4 conjugate gradient iterations, even close to the turning point.
210
VII Least-Squares Solution of Nonlinear Problems Figure 3.1
7
u, (0.5, 0.5) 8.00
I i i i i i i i i i
4.80
Figure 3.2
6.40 8.00 LAMBDA
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
211
In Fig. 3.2 we show the maximal value (reached at xt = x2 = 0.5) of the computed solution uh as a function of X; the computed turning point is at X = 6.8591... . The initialization of the conjugate gradient algorithm used to solve system (3.62), (3.64), via least squares, was performed according to Remark 3.4. 3.5.5. Further comments
The least-squares conjugate gradient continuation method described in Sees. 3.5.2, 3.5.3, and 3.5.4 has been applied to the solution of nonlinear problems more complicated than (3.70); among them let us mention the Navier-Stokes equations for incompressible viscous fluids at high Reynold's number and also problems involving genuine bifurcation phenomena like the Von Karman equations6 for plates. The details of these calculations can be found in Reinhart [1], [2] and Glowinski, Keller, and Reinhart [1].
4. Transonic Flow Calculations by Least-Squares and Finite Element Methods
4.1. Introduction
In Chapter IV, Sec. 3, we considered the nonlinear elliptic equation modelling the subsonic potential flows of an inviscid compressible fluid. In this section, which closely follows Bristeau, Glowinski, Periaux, Perrier, Pironneau, and Poirier7 [2] (see also G.L.T. [3, Appendix 4]). we would like to show that the least-squares conjugate gradient methods of Sec. 2 and 3 can be applied (via convenient finite element approximations) to the computation of transonic flows for similar fluids. Given the importance and complexity of the above problem, we would like to point out that the following considerations are just an introduction to a rather difficult subject. Many methods, using very different approaches, exist in the specialized literature, and we shall concentrate on a few of them only (see the following references for other methods). We would also like to mention that from a mathematical point of view, the methods to be described in the following sections are widely heuristical. 6 7
For which we refer to the monograph by Ciarlet and Rabier [1] (and the references therein). B.G.4P. in the sequel
212
VII Least-Squares Solution of Nonlinear Problems
4.2. Generalities. The physical problem The theoretical and numerical studies of transonic potential flows for inviscid compressible fluids have always been very important questions. But these problems have become even more important in recent years in relation to the design and development of large subsonic economical aircrafts. From the theoretical point of view, many open questions still remain, with their counterparts in numerical methodology. The difficulties are quite considerable for the following reasons: (1) The equations governing these flows are nonlinear and of changing type (elliptic in the subsonic part of the flow; hyperbolic in the supersonic part). (2) Shocks may exist in these flows corresponding to discontinuities of velocity, pressure, and density. (3) An entropy condition has to be included "somewhere" in order to eliminate rarefaction shocks, since they correspond to nonphysical situations. Concerning the fluids and flows under consideration, we suppose that these fluids are compressible and inviscid (nonviscous) and that their flows are potential (and therefore quasi-isentropic) with weak shocks only; in fact this potential property is no longer true after a shock (cf. Landau and Lifchitz [1]). In the case of flows past bodies, we shall suppose that these bodies are sufficiently thin and parallel to the main flow in order not to create a wake in the outflow. 4.3. Mathematical formulation of the transonic flow problem. References 4.3.1. Governing equations If Q is the region of the flow and T its boundary, it follows from Landau and Lifchitz [1] that the flow is governed by the so-called full potential equation: V • pu = 0 in Q,
(4.1)
where 1 - - [(y
+
i y
^ _
1)]c2
u = \(j>, and (a) (b) (c) (d)
1
,
(4.2) (4.3)
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
213
Figure 4.1
4.3.2. Boundary conditions
For an airfoil B (see Fig. 4.1), the flow is supposed to be uniform on F^, and tangential at TB. We then have (4.4)
— ^ - = Oonr B .
(4.5)
Since only Neumann boundary conditions are involved in the above case, the potential
(4.6)
4.3.3. Lifting airfoils and the Kutta-Joukowsky condition For two-dimensional flows a slit E (see Fig. 4.2) has to be introduced in Q in order to render the potential (j) single valued, and we choose (j) to be discontinuous across X. The circulation J3 is another unknown depending on the
Figure 4.2
UK
214
VII Least-Squares Solution of Nonlinear Problems
boundary conditions and the geometry of B. Along S, the following relation is required:
p
= p~ at T.E.
(4.7)
condition (p: pressure),
(4.8)
which, by applying the Bernouilli law, may also be written as \\(j)+\ = | V 0 - | a t T . E . .
(4.9)
Observe that (4.9) is a nonlinear relation. Remark 4.1. In two-dimensional cases we can use linear formulations of the Kutta-Joukowsky condition. Furthermore, if there is no cusp at T.E., it has been proved in Ciavaldini, Pogu, and Tournemine [3], for strictly subsonic flows (i.e., | V> | < C,,, everywhere), that the physical solution is characterized by \cj) = 0 at T.E,
(4.10)
which can be taken as the Kutta-Joukowsky condition. A similar result has not yet been proved for genuine transonic flows. Let us again emphasize that (4.10) is no longer true if there is a cusp at T.E. (as is the case, for example, for the celebrated Korrfs airfoil). The treatment of the Kutta-Joukowsky condition for three-dimensional flows is much more complicated; we refer to B.G. 4P. [1] and Bristeau, Glowinski, Perrier, Periaux, and Pironneau [1] for the practical implementation of the three-dimensional Kutta-Joukowsky condition. 4.3.4. Shock conditions Across a shock the flow has to satisfy the Rankine-Hugoniot conditions (pu-n) + = ( p u - n ) _
(4.11)
(where n is normal at the shock line or surface); the tangential component of the velocity is continuous.
(4.12)
A suitable weak formulation of (4.1)-(4.3) will take (4.11), (4.12) into account automatically. 4.3.5. Entropy condition This condition can be formulated as follows (see Landau and Lifchitz [1] for further details): Following the flow we cannot have a positive variation of velocity through a shock, since this would imply a negative variation of entropy which is a nonphysical phenomenon. The numerical implementation of (4.13) will be discussed in Sec. 4.6.
(4.13)
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
215
4.3.6. Some references: Synopsis
The mathematical analysis of the above transonic flow problem is quite difficult. Some standard references are Bers [1] and Moravetz [l]-[5] (see also Landau and Lifchitz [1] and Courant and Friedrichs [1] for the physical aspects). From the numerical point of view, the more commonly used finite difference methods have originated from Murman and Cole [1], and we shall mention, among many other references, Bauer and Garabedian, and Korn [1], Bauer, Garabedian, Korn, and Jameson [1], [2], Jameson [1], [2] ,[3], [4], Hoist [1], and Osher [1], and the bibliographies therein (see also Hewitt and Hillingworth and co-editors [1]). The above numerical methods use the key idea of Murman and Cole, which consists of using a finite difference scheme (centered in the subsonic part of the flow) backward8 (in the direction of the flow) in the supersonic part. The switching between these two schemes is automatically accomplished via a truncation operator active only in the supersonic part of the flow (see Jameson and Hoist, loc. cit. for more details). Relaxation or alternatingdirection methods (or a combination of both) are then used to solve the resulting nonlinear system. These finite difference methods of solution have been extended to finite elements (of quadrilateral type) by Eberle [1], [2], Deconinck and Hirsh [1], and Amara, Joly, and Thomas [1]. The methods to be described in Sees. 4.4, 4.5, and 4.6 allow the use of triangular (or tetrahedral) elements and are well suited to a least-squares conjugate gradient solution. 4.4. Least-squares formulation of the continuous problem
In this section we do not consider the practical implementation of (4.13); we only discuss the variational formulation of (4.1)-(4.5), (4.11), (4.12) and of an associated nonlinear least-squares formulation. 4.4.1. A variational formulation of the continuity equation
For simplicity we consider the situation of Fig. 4.3 which shows a symmetric flow, subsonic at infinity, around a symmetric airfoil; thus the KuttaJoukowsky condition is automatically satisfied. For practicality (but other approaches are possible) we imbed the airfoil in a "large" bounded domain. Using the notation of Sec. 4.3, the continuity equation and the boundary conditions are = 0 in Q,
' One also says upwinded or one sided.
(4.14)
216
VII Least-Squares Solution of Nonlinear Problems Figure 4.3
with 1/(7-1)
(4.15)
= Po( 1 ~ and
d
(4.16)
on F (= F B u r^), we define g by gr = Oonr B ,
g= ^u^n^onr^.
(4.17)
We clearly have dd>
f
p — = g on F and
\ g dT = 0.
(4.18)
An equivalent variational formulation of (4.14), (4.18) is (4.19)
p(
where (cf. Adams [1] and Necas [1]), for p > 1, W1'P(Q) is the Sobolev space defined by = \v\ve
,V
(with H^Q) = WU2((J)); the function 0 is determinated only to within an arbitrary constant. Remark 4.2. The space W1' °°(Q) (space of Lipschitz continuous functions) is a natural choice for (p since physical flows require (among other properties) a positive density p; therefore, from (4.2), (4.3), cp must satisfy + l\ 1 / 2 —-1 C,a.e.onQ.
(4.20)
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
217
4.4.2. A least-squares formulation of (4.19) For a genuine transonic flow, problem (4.19) is not equivalent to a standard problem of the calculus of variations (as is the case for purely subsonic flows; see Chapter IV, Sec. 3). To remedy this situation, and—in some sense—convexify the problem under consideration, we consider, as in Sec. 3.3 of this chapter, a nonlinear least-squares formulation of the transonic flow problem (4.18), defined as follows: Let X be a set of feasible transonic flow solutions; the least-squares problem is then Min ./(£),
(4.21).
with \ \
2
d x ,
(4.22)
where, in (4.22), y(£) ( = y) is the solution of the state equation: Find y e H\Q)/U such that
[\y-Svdx
= f p(£)V£ -Vvdx-
\ gvdr,
V u e H\Q).
(4.23)
If the transonic flow problem has solutions, these solutions solve the leastsquares problem and give the value zero to the cost function J.
4.5. Finite element approximation and least-squares-conjugate gradient solution of the approximate problems We consider only two-dimensional problems; but the following methods can be (and have been) applied to three-dimensional problems. 4.5.1. Finite element approximation of the nonlinear variational equation (4.19) We still consider the nonlifting situation of Sec. 4.4.1. Once the flow region has been imbedded in a large bounded domain Q, we approximate this later domain by a polygonal domain Q h ; with ^ a standard triangulation of Q h , we approximate H^O) (and in fact WUp(Q), V p > 1) by Hi = {vh\vh6 C°(Q,), vh\TePu
V T e 3Th),
(4.24)
where, in (4.24), Px is the space of the polynomials in two variables of degree < 1. We prescribe the value zero for the potential at T.E.; this leads to .) = 0}.
(4.25)
218
VII Least-Squares Solution of Nonlinear Problems
We clearly have dim Hj, = dim Vh + 1 = number of vertices of 2Th.
(4.26)
We then approximate the variational equation (4.19) by: Find (j)h e Vh such that f P(
VvheVh
(4.27)
where, in (4.27), gh is an approximation of the function g of (4.17) (and Th = dQh). The above discrete variational formulation implies that p d^/dn\T = g is approximately satisfied automatically. Let SSh = {w,-}^! be a vector basis of Vh (with Nh = dim Vh); then (4.27) is equivalent to the nonlinear finite-dimensional system
=t
tjwj, f P(^
ghWi dY,
Mi =
(4.28) j}7= i is the set of the vertices of Fh different from T.E., we take (for @h) the set defined by «»={w^i
with W j .eF A , l,
Wj(Pk) = 0,
Vj=l,...,Nm, Vfc^j;
we then have (^ = 4>h(Pj). From the above choice for H,| and Vh, there is no problem of numerical integration since, in (4.27), (4.28), \<j)h, \vh (and therefore p(>ft)) are piecewise constant functions. 4.5.2. Least-squares formulation of the discrete problem (4.27), (4.28) For simplicity we set Qh = Q, Th = F. Combining the results of Sees. 4.4.2 and 4.5.1, we introduce the following least-squares formulation of the approximate problem (4.27), (4.28): Min Jh(Q,
(4.30)
ZheXh
where, in (4.30), Xh is the set of the feasible discrete solutions and \\^yh{U)\2dx,
(4.31)
with yh(^h) ( = yh) the solution of the discrete variational state equation: Find yh e Vh such that f \yh • \vh dx = f p(^)\^h
• \vh dx~ [gh vh dY,
V vh e Vh.
(4.32)
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
219
4.5.3. Conjugate gradient solution of the least-squares problem (4.30)-(4.32) We follow B.G. 4P. [1], [2], Bristeau, Glowinski, Periaux, Perrier, and Pironneau [1], and Periaux [2]; a preconditioned conjugate gradient algorithm for solving (4.30)-(4.32) (with Xh = Vh) is as follows. Step 0: Initialization
(4.33)
glfrom g°h £ K, f \g°h • \vh dx = (J'h(
V ^ e Vh,
(4.34)
and set z°H=g°H.
(4-35)
Then for n > 0, assuming that ^>nh,gnh, z\ are known, compute 4>1, + l, gl+1, zJJ+1 by the following. Step 1: Descent. Compute A" = Arg min Jh(4>nh - W),
(4.36)
^+1 = ^-A%".
(4.37) +l
Step 2: Construction of the new descent direction. Define gl gl+1eVh,\
Jn
\g"h+1 • Vvh dx = < J ^ + ' ) , vh),
by
VvheVh;
(4.38)
then
znh+1 = g"h+1 + y n + 1 z n h ,
(4.40)
n = n + 1 go to (4.36). The two nontrivial steps of algorithm (4.33)-(4.40) are as follows: (i) The solution of the single-variable minimization problem (4.36); the corresponding line search can be achieved by the methods mentioned in Sec. 3.4. We observe that each evaluation of Jh(ih), for a given argument £h, requires the solution of the linear approximate Neumann's problem (4.32) to obtain the corresponding yh. (ii) The calculation of g"h+l from
220
VII Least-Squares Solution of Nonlinear Problems
Calculation ofJ'h(4>l) and gfJJ: Due to the importance of step (ii), let us describe the calculation of J'h((f>1,) and g\ in detail (for simplicity we suppose that p0 = 1). By differentiation we have \x, ^n where dyh is from (4.32), the solution of: Syhe Vh and V vheVh,we have
(4.41)
dx + V£h • Vvh dx. (4.42) \ p Jn Jn Since P&) =-- (1 - K V&|2 )x, with K = )[(7 - 1)/(y + 1)] and a = 1/(7- 1), we have •\v
Jn
tdx=
(Vet
(4.43) From (4.41)^(4.43) it follows that
ff
dx. (4.44)
Jn Finally J'h(^h) can be identified with the linear functional on Vh defined by f /*&Vyk • ^ dx-2Ka\ (p(ih))2 " J V ^ • \yh \£h • \rjh dx. Jn Jn It is then quite easy to obtain g"h+1 from
(4.45)
Remark 4.3. An efficient discrete Poisson's solver will be a basic tool if one uses the above conjugate gradient algorithm (4.33)-(4.40). 4.1. Show that algorithm (4.33)-(4.40) is a particular case of algorithm (2.13)-(2.19), (2.21) of Sec. 2.2. Identify the matrix S.
EXERCISE
4.5.4. Numerical solution of a test problem. We apply the above iterative and approximation methods to the numerical simulation of a flow around a disk; we suppose that the flow is subsonic and uniform at infinity. From the symmetry, the Kutta-Joukowsky condition is automatically satisfied. If | u „ | is sufficiently small, the flow is purely subsonic, and a very good solution is obtained in very few iterations of algorithm (4.33)-(4.40) ( ~ 5 iterations). For greater values of lu^l, a supersonic pocket appears, and if we plot the computed 9 Mach distribution on the skin of the 9
We still have a very fast convergence (~ 10 iterations).
221
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods Figure 4.4
_ , . Expansion shock M<1
Figure 4.5
Physical shock M< 1
Physical shock M<1
M <1
body, we observe the distribution of Fig. 4.4, showing a rarefaction (or expansion) shock which is a nonphysical phenomenon. The correct Mach distribution is shown on Fig. 4.5. From the above numerical experiments, it appears that in our numerical process we have to include a device (some dissipation, for example) which eliminates those nonphysical shocks violating the entropy condition (4.13). The numerical implementation of (4.13) is discussed in Sec. 4.6. 4.6. Numerical implementation of the Entropy Condition
4.6.1. Generalities: Synopsis Several methods based on penalty and/or artificial viscosity have been discussed and numerically tested in B.G.4P. [1], Bristeau [2], Periaux [2], Bristeau, Glowinski, Periaux, Perrier, and Pironneau [1], and Poirier [1]. In Sec. 4.6.2. we describe (following B. G. 4P. [2]) an interior penalty method, and in Sec. 4.6.3 we describe a method based on upwinding of the density (as in Eberle [1], [2], Deconinck and Hirsch [1], and Hoist [1]).
222
VII Least-Squares Solution of Nonlinear Problems
Figure 4.6
If we go back to the example of Sec. 6.5.4 and try to represent the corresponding least-squares functional £h -> Jh(^h), it is very likely that it looks like the graph of Fig. 4.6 (if we suppose | u^ | sufficiently large). Our feeling about Fig. 4.6 is motivated by the fact that lh is a very powerful attractor for almost any iterative method, particularly for algorithm (4.33)(4.40); on the other hand, the physical solution lacks these attractor properties (except possibly in a very small neighborhood of this solution). From these observations a reasonable idea is to replace the continuous curve in Fig. 4.6 with the discontinuous one, corresponding to a functional t,h -» J*(£h), taking a very large value at the nonphysical solution \h, and taking its minimal value close to
(4.46)
where in (4.46), [
(4.47)
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
223
Figure 4.7
combining (4.46), (4.47), we shall use the entropy condition
M 3 < 0.
(4.48)
This choice is motivated by the fact that our finite element method gives us direct access to velocity variables. 4.6.2.2. Finite element realization o/(4.48) by an interior penalty method. Let us consider two adjacent triangles, as in Fig. 4.7 [actually the following method can be (and has been) applied to three-dimensional problems]; we denote by n the unit vector of the normal at their common edge AB, oriented from Tt to T2. Since we use a piecewise linear approximation, we have u,,^ = V
7=1,2
(4.49)
with \cj)h piecewise constant. Moreover, since we use C°-conforming elements, the component along AB of the velocity nh = \(f>h is continuous, unlike the normal component which is clearly discontinuous; from these considerations the discontinuities of nh are supported by the local normal components of the velocity. From these observations and from (4.49), we introduce the functional Eh defined as follows:
where the summation is done with respect to the common edges of the adjacent triangles of J^, where cr is a coefficient which takes the local size of the triangulation into account, and where t+ = max(0, t),
VtelR
([d(j)h/dni'] denotes the jump of normal derivative in the direction of n,). Remark 4.4. We observe that in fact (4.50) is very close to the integral over Q of the sixth power of a truncated second-order derivative; for homogeneity this suggests that we take a, = dl (4.51)
224
VII Least-Squares Solution of Nonlinear Problems
where d( is the length of the common edge and <x = — 4 (we should take a = —5 for one-dimensional problems and a = — 3 for three-dimensional problems). Combining the functional Eh and the approximate problem (4.30)-(4.32) of Sec. 4.5.2, we obtain the following discrete problem, containing a dissipative device in order to avoid nonphysical shocks: ^,
(4.52)
where Xh is the set of the feasible discrete solutions and where k) I2 dx + rEh{^\
(4-53)
with r > 0 and yh(^h)( = yh), the solution of the discrete variational state equation: Find yh e Vh such that
Ja
Jr
ghvhdT,
VvkeVh.
(4.54)
To solve (4.52)-(4.54), we can use a variant of the conjugate gradient algorithm (4.33)-(4.4O) (with Jh replaced by Jrh). The numerical results obtained by the above methods are quite good since the sonic transition (without shock) from the subsonic to the supersonic region is very well approximated. Furthermore, the shocks are well located and are very sharp. We refer to Periaux [2] and also to Sec. 4.7 where numerical results obtained by the above methods are presented and discussed. A practical and very important problem is the proper choice of r, since for a given airfoil the optimal r seems to be a complicated and sensitive function of | ux | and of the angle of attack; in Sec. 4.6.2.3 we shall discuss a technique which seems to be effective for the removal (or at least the reduction) of sensitivity to the choice of r. Remark 4.5. We complete Remark 4.4; since Eh(£h) appears as the integral over Q of the sixth power of a discrete truncated second-order derivative of £h, we can associate to Eh—by differentiation—a discrete fourth-order nonlinear operator. Since Eh is a convex functional, its differential is a monotone operator. From these properties the addition of rEh to Jh may be interpreted as a nonlinear fourth-order elliptic regularization process. Remark 4.6. We justify the title of Sec. 4.6.2 by observing that in fact the functional Eh introduced to regularize our problem is a variant (more complicated) of the interior penalty functionals discussed in Douglas and Dupont [1] and Wheeler [1].
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
225
Remark 4.7. We have performed computations with Eh of (4.50) using exponents different from 2 and 3; actually these two exponents seem to be the optimal combination in view of the quality of the computed solutions. EXERCISE
4.2. Compute the differential of the functional Jrh denned by (4.53).
4.6.2.3. A nonlinearly weighted interior penalty method. From the comments in Sec. 4.6.2.2 about the sensitivity to the choice of r in Jrh—which is clearly related to the strong nonlinearity of our problem—it is very tempting to introduce a local control of the regularization process associated with ([_d(j)h/dn]+)6. To achieve such a goal, we have introduced, in the above functional Eh (defined by (4.50)), a nonlinear weight directly related to the local values of the density; more precisely, instead of using Eh, we use Eh defined by (4.55) where 1/pf is a symbolic notation denned by either
<456)i
or 1 _ 1 / 1 Pi
2" \Pn
1 Pa
with
,-U
\™hl
V
(4 57)
where
Tn, Ti2 being the adjacent triangles of 2Th having the ith edge in common. Remark 4.8. From our computer experiments, a " good " value for n in (4.55) is n = 2. Replacing Eh by Eh, we obtain a variant of the approximate problem (4.52)(4.54) which can be solved by the same type of iterative methods; computer experiments performed with Eh have shown that for a NACA 0012 airfoil, the same r was optimal (or nearly optimal) in the conditions listed in Table 4.1. The corresponding results are described in Sec. 4.7.
226
VII Least-Squares Solution of Nonlinear Problems Table 4.1 Angle of Attack (Degrees)
Mach Number at Infinity ( M J
6 1 0 0
0.6 0.78 0.8 0.85
4.6.2.4. An interior penalty method using density jumps. The interior penalty method discussed in Sec. 4.6.2.3 combines two effects: (i) It penalizes rarefaction shocks via [u] +. (ii) Because of the nonlinear weight that we have introduced, the regularization effect is reinforced in regions at high Mach number, since in these regions p is "small". It is then natural to look for a variant of (4.55), combining both effects more closely; we start from the observation that in a physical shock we have (still following the stream lines)
rn : 0.
(4.58)
This jump condition (4.58) leads to the following variant of the entropy functionals Eh and Eh of Sees. 4.6.2.2 and 4.6.2.3 (whose notation has been retained):
In (4.59) the notation is self-explanatory, except for vv; which is defined by
the indices; = 1, 2 corresponding to the two adjacent triangles of &~h having the ith edge in common. It is then quite easy to formulate an approximate problem in which Eh or Eh is replaced by Rh; moreover, the same type of iterative methods will hold for this new type of approximate problem. Numerical experiments have to be performed to check the validity of this new approach and also to make an optimal choice for the two exponents /? and n in (4.59). 4.6.3. Implementation of the entropy condition by upwinding of the density 4.6.3.1. Generalities: Synopsis. Density upwinding has been extensively used (see Eberle [1], [2], Hoist [1], and Deconinck and Hirsh [1] and the references
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
227
Figure 4.8
therein) in order to eliminate nonphysical shocks; these upwinding techniques have been very effective, coupled with alternating-direction methods (implicit or semi-implicit; see Hoist [1] and Deconinck and Hirsch [1]) if the computational mesh is regular (finite differences or regular finite element grids). In particular, their application, combined with finite element techniques, have been limited (see Eberle [1], [2] and Deconinck and Hirsh [1]) to quadrilateral elements on quasiregular quadrangulations (fairly close to finite difference methods, in our opinion). In Sec. 4.6.3.2 we would like to discuss a method (due to M. O. Bristeau) which also makes use of an upwinding of the density; this method can be used with simplicial10 finite elements (triangles in two dimension, tetrahedra in three dimensions) and has been very effective for computing flows at high Mach numbers and around complicated two- and three-dimensional geometries. 4.6.3.2. A modified discrete continuity equation by upwinding of the density in the flow direction. Following Jameson [l]-[4] and Bristeau [2], [3], we may write the continuity equation (4.14) in a system of local coordinates {s, n}, where (see Fig. 4.8) for a two-dimensional flow, s is the unit vector of the stream direction (i.e., s = u/1 u | if u # 0) and n is the corresponding normal unit vector (conventionally oriented). Using {s, n} and setting11
we obtain (from (4.14))
10 1
We use here the terminology of Ciarlet [1], [2]. ' U is the Mach number M*.
228
VII Least-Squares Solution of Nonlinear Problems
the elliptic-hyperbolic aspect of the problem is clear from (4.61). Actually (4.61) can also be written
0
+
VV O
(462)
*
[we have 1/2/ca = (y + l)/2]. EXERCISE
4.3. Prove (4.61), (4.62).
We use (4.62) to modify the discrete continuity equation (4.27) as follows: Find 4>h e Vh such that
Vvh dx + ^- £ (^ hs{U2h - l)+\
dx VvheVh.
(4.63)
The approximate problem (4.63) has been introduced by M. O. Bristeau and is a finite element variant of a finite difference scheme due to Hoist [1]. In (4.63): (i) hs is a measure of the local size of the finite element mesh in the flow direction, (ii) (d/ds)h is an approximation of ( V ^ / | V(/>A|) • V, (iii) Uh, ph are upwinded approximations of U and p, respectively. More precisely, we write the second integral in the left-hand side of (4.63) as follows.
= (-1) Z ^ Z ^ E m(T)hs(T)(U2h - 1 ) + V ^ - V ^ - ^ , (4.64) where (a) l,h = {P^o i s the set of the vertices of STh, with Po = T.E.; (b) W; is the basis function of HI (cf. (4.24)) associated with Pt by w, e Hi
Vi = 0,...,Nh,
Wi(Pd
= 1,
WiiPj) = 0,
\/jjt
i; (4.65)
(c) ^ is the subset of 2Th consisting of those triangles having Pi as a common vertex; (d) m(T) = meas(T);
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
229
Figure 4.9
(e) {iiji}j= i is an approximation of \(/)J| \
I (4.66)
'50»
(others averaging methods are possible); (f) hs(T) is the length of T in the s direction, i.e.,
K(T) =
(4.67) v
kT
=i
ds
where wfcr, k = 1, 2, 3 are the basis functions associated with the vertices PkT of T; actually (4.67) can also be written (4.68)
(g) We define C/j as follows: With each vertex Pt o{STh we associate an inflow triangle T{ which is the triangle of ?7~h having Pt as vertex and which is crossed by the vector u ; — iuij}j=I pointing to Pt (as shown on Fig. 4.9); we then define Uf as
Uf =
cl
230
VII Least-Squares Solution of Nonlinear Problems
and for each triangle T of 3~h, dwkT
ds
k=l
3
T
(4.69)
dwkT
I
ds
k=l
T
(h) We finally obtain ph e Hi as follows: As for Uf, we define pt by Pi =
P(
and then (4.70)
Ph =
4.6.3.3. Some brief comments on the least-squares sohrtion of (4.63). For solving the discrete upwinded continuity equation (4.63), we can use the following leastsquares formulation: (4.71) where
with yh(i;h) (=yh), the solution of the following linear variational equation: Find yh e Vh such that
f
f Vj>h • \vh dx =
Jn
+
f p((j)h)S/4>h • Vvh dx -
Jn
\ ghvh dT Jr
2iL I (Is W* "1}+V>"' Vp "), t; ' 1 JX'
V £ Fft (472)
"* '
Since J h is a nondifferentiable functional of £,,, to solve (4.71) we have used (instead of algorithm (4.33)-(4.40)) a generalization of the conjugate gradient method due to Lemarechal [1], [2] which also applies to the minimization of nondifferentiable functional (actually good results are also obtained if one uses algorithm (4.33)-(4.40) with Jh( •) denned as in (4.71); however, more iterations are needed). 4.7. Numerical experiments In this section we shall present some of the numerical results obtained using the above methods. The results of Sec. 4.7.1 are related to a NACA 0012 airfoil; those of Sec. 4.7.2 (resp., 4.7.3) to a Korn airfoil (resp., a two-piece airfoil).
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
231
4.7.1. Simulation of flows around a NACA 0012 airfoil. As a first example we have considered flows around a NACA 0012 airfoil at various angles of attack and Mach numbers at infinity. The corresponding pressure distributions on the skin of the airfoil are shown on Figs. 4.10-4.17 in which the isomach lines (in the supersonic region only on Figs. 4.10-4.14) are also shown. The results shown in Figs. 4.10-4.14 have been obtained using the interior penalty method of Sec. 4.6.2; those of Fig. 4.15-4.17 have been obtained using the upwinding method of Sec. 4.6.3. We observe that the physical shocks are quite neat and also that the transition (without shock) from the subsonic to the supersonic region is smoothly restituted, implying that the entropy condition has been satisfied. The above numerical results are very close to those obtained by various authors using finite difference methods (see, particularly, Jameson [1]). 4.7.2. Flow around a Horn's airfoil
In Fig. 4.18 we have represented the pressure distribution corresponding to the flow around a Korn's airfoil atM 0O = 0.75 and a. — 0.11; the computation method is the interior penalty method of Sec. 4.6.2. The agreement with a finite difference solution is good, as indicated in Fig. 4.18. 4.7.3. Flows around a two-piece airfoil
The tested two-piece airfoil is shown on Figs. 4.19 and 4.20. Each piece is a NACA 0012 airfoil (the body No. 1 is the upper body). The pressure distribution and the isomach lines (computed by the interior penalty method of Sec. 4.6.2) are shown on Figs. 4.19 and 4.20. We observe that the region between the two airfoils acts as a nozzle; we also observe supersonic regions, in particular, between the two airfoils.
4.8. Transonic flow simulations on large bounded computational domains
Consider the situation depicted in Fig. 4.3; if the supersonic zone extends far from the airfoil, it is necessary to use a very large computational domain. Let 0 be the origin of coordinates; it is then reasonable to take the circle {x eU2,r = Rx} for r ^ , with r = y/x{ + xf. Now suppose that the disk of center 0 and radius Ro is sufficiently large to contain the airfoil B in its interior; we then introduce Qi = {x e U2, 0 < r < Ro, x $ B}, Q2 = {xeR2,J?0
Figure 4.10. NACA 0012 airfoil. Ma = 0.6; a = 6. (Interior penalty method.)
5o
o
Figure 4.11. NACA 0012 airfoil. M» = 0.78; a = 1. (Interior penalty method.)
-1
tr oa
tn
§'
5'
9
o
3
234
VII Least-Squares Solution of Nonlinear Problems
B o
<
U <
z
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
235
-II
•g
<
Figure 4.14. NACA 0012 airfoil. M . = 0.85; a = 0. (Interior penalty method.)
-1
5 o
3
Z o
I
Figure 4.15. NACA 0012 airfoil. Mx = 0.6; a = 6. (Computation with upwinding of the density.)
-1
Pi
Tl
a
c
VI
O
VII Least-Squares Solution of Nonlinear Problems
238 CP4
.25
.5
.75
Figure 4.16. NACA 0012 airfoil. Mm = 0.85;a = 0. (Computation with upwinding of the density.)
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
239
-1
.25
.5
.75
Figure 4.17. NACA 0012 airfoil. M . = 0.9; a = 0. (Computation with upwinding of the density.)
Figure 4.18. Korn airfoil. Ma — 0.75; a = 0.11. (Interior penalty method.)
1.0
Finite difference ^— (Jameson)
o
o
Z
c
Figure 4.19. Bi NACA 0012 airfoil. M«, = 0.6; a = 0. (Interior penalty method.)
lower surface 2 upper surface 1
lower surface 1 upper surface 2
W
5'
c
9
o
Figure 4.20. Bi NACA 0012 airfoil. Mm = 0.6; a = 6. (Interior penalty method.)
. upper surface 2
o
o
C
C/5
4 Transonic Flow Calculations by Least-Squares and Finite Element Methods
243
With {r, 9} as a standard polar coordinate system associated with 0, we define the following new variables: — X2,
V X — {X l 5 X;
(4.73) Z2 = 6,
V{r,0}eQ 2 ;
we use the notation £, = {^1; £2} in the sequel. We now introduce Q t = £lu and f / R \ 1 Q 2 = « e U 2 , R 0 < ^ < R O ( 1 + L o gR- ^ , 0 < ^ 2 < 2n};
I
\
o/
J
we then define
f
H 1 =
COS
<^2> ^o s n l ii)
= v2(R0, £2), V l2 e ] 0 , 2TT[, ^ ( ^ J , 0) = C 2 (^, 2?r),
The space H 1 is clearly isomorphic to H 1(Q.) via the transformation v -+ v: H -+H1(Q): 1
v1(x1,x2)
ifx = {x1;
u(x) =
(4.74)
v2\R0n+Log~\e Using the above properties, the continuity equation (4.19) is now formulated as follows:
Find | e f l ' such that V P E H 1
2lc
(4.75) with {«loo>«2oo} = "ao
244
VII Least-Squares Solution of Nonlinear Problems
and dS
^i2 +
R-2
0
2
\
~| i/(7-D
^2
\e2(R0- S,)/Ro
Kl
I
J
' (4.76) The discretization of (4.75), (4.76) can be done as follows: On the one hand, one uses in Q.1 (which is part of the physical space) a standard finite element approximation taking into account the possible complexities of the geometry; on the other hand, since fi2 is a rectangle, one may use a finite difference discretization which can in fact be obtained via a finite element approximation on a uniform triangle or quadrilateral grid (like the one in Fig. 3.1 of Sec. 3.5.4 of this chapter). Remark 4.9. From the above decomposition of the computational domain (in Cll and Q2), it is natural to solve the approximate problems by iterative methods, taking this decomposition into account, and also the fact that finite differences can be used in Q2 (since finite differences allow special solvers on Q2). We are presently working on such methods and also on their generalization to three-dimensional problems; the corresponding results will be presented in a forthcoming paper.
5. Numerical Solution of the Navier-Stokes Equations for Incompressible Viscous Fluids by Least-Squares and Finite Element Methods
5.1. Introduction. Synopsis
The numerical solution of the Navier-Stokes equations for incompressible viscous fluids has motivated so many authors that giving a complete bibliography has become an impossible task. Therefore, restricting our attention to very recent contributions making use of finite element approximations, we shall mention, among many others, B. G. 4P. [1], Bristeau, Glowinski, Periaux, Perrier, and Pironneau [1], Bristeau, Glowinski, Mantel, Periaux, Perrier, and Pironneau [1], Glowinski, Mantel, Periaux, Perrier, and Pironneau [1], Gartling and Becker [1], [2], Hughes, Liu, and Brooks [1], Temam [1], Bercovier and Engelman [1], Fortin and Thomasset [1], Girault and Raviart [1], Le Tallec [2], [3], Johnson [2], Glowinski and Pironneau [1], [2], Gresho, Lee, Chan, and Sani [1], Rannacher [1], Benque, Ibler, Keramsi, and Labadie [1], Thomasset [1], and Brooks and Hughes [1]; see also the references therein.
5 Numerical Solution of the Navier-Stokes Equations
245
In this section (which very closely follows Bristeau, Glowinski, Mantel, Periaux, Perrier, and Pironneau [1]) we would like to discuss several methods for the effective solution of the above Navier-Stokes problems in the steady12 and unsteady13 cases. The basic ingredients of the methods to be described are the following: (i) Mixed finite element approximations of a pressure-velocity formulation of the original problem. (ii) Time discretizations of the unsteady problem by finite differences; several schemes will be presented. (iii) Iterative solution of the approximate problems by least-squares conjugate gradient methods (possibly combined with an alternatingdirection method). (iv) Efficient solvers for the discrete Stokes problems. The possibilities of the above methodology will be illustrated by the results of various numerical experiments concerning nontrivial two-dimensional flows.
5.2. Formulation of the steady and unsteady Navier-Stokes equations for incompressible viscous fluids Let us consider a Newtonian, viscous, and incompressible fluid. If Q and F denote the region of the flow14 and its boundary, respectively, then this flow is governed by the Navier-Stokes equations $ - vAu + (u • V)u + Sp = f in Q,
(5.1)
ot \ • u = 0 in Q (incompressibility condition),
(5.2)
which in the steady case reduce to - vAu + (u • V)u + \p = f in Q,
(5.3)
V • u = 0 in Q.
(5.4)
(i) u = {wJfL j is the flow velocity, (ii) p is the pressure, (iii) v is the viscosity of the fluid (v = I/Re, where Re is the Reynold's number), (iv) f is the density of external forces;
12 13 14
One also says stationary. One also says nonstationary. Q a W, N = 2, 3 in practice.
246
VII Least-Squares Solution of Nonlinear Problems
in (5.1), (5.3), (u • V)u is a symbolic notation for the nonlinear (vector) term:
Boundary conditions have to be added; for example, in the case of the airfoil, B of Fig. 4.1 of Sec. 4.3.2 of this chapter, we have (since the fluid is viscous) the following adherence condition: u = 0ondB = rB.
(5.5)
Typical conditions at infinity are u=
Uoo ,
(5.6)
where u^ is a constant vector (with regard to the space variables, at least). Finally, for the time-dependent problem (5.1), (5.2), an initial condition such as u(x, 0) = uo(x) a.e. on Q,
(5.7)
where u0 is given, is usually prescribed. Other boundary and/or initial conditions may be prescribed (periodicity in space and/or time, pressure given on 3Q or on a part of it, etc.). In two dimensions it may be convenient to formulate the Navier-Stokes equations using a stream-function-vorticity formulation (see, e.g., Bristeau, Glowinski, Periaux, Perrier, and Pironneau [1, Sec. 4], Fortin and Thomasset [1], Girault and Raviart [1], Glowinski and Pironneau [1], Reinhart [1], and Glowinski, Keller, and Reinhart [1]). To conclude this section, let us mention that a mathematical analysis of the Navier-Stokes equations for incompressible viscous fluids can be found in, e.g., Lions [1], Ladyshenskaya [1], Temam [1], and Tartar [1]. 5.3. A mixed finite element method for the Stokes and Navier-Stokes problems 5.3.1. Synopsis In this section we discuss a mixed finite element approximation of the NavierStokes problems which have been introduced in Sec. 5.2. For simplicity we shall begin our discussion with the approximation of the steady Stokes problem for incompressible viscous fluids, i.e., - vAu + \p = f in Q, V • u = 0 in Q;
(5-8)
as boundary conditions, we choose u = g on T
(5.9)
(with j r g • n dF = 0, n being the unit vector of the outward normal at F), more general boundary conditions are discussed in Appendix 3.
5 Numerical Solution of the Navier-Stokes Equations
247
Also, for simplicity, in the following we suppose that Q is a bounded polygonal domain of U2; but the following methods are easily extended to domains with curved boundary in U2 and U3. 5.3.2. A mixed variational formulation of the Stokes problem (5.8), (5.9) 5.3.2.1. Some functional spaces: Standard formulation of the Stokes problem. The following (Sobolev) spaces play an important role in the sequel (in fact they have been extensively used in the other parts of this book):
= WHa) sH\Cl),
= {(j) e Hl(Q), ^
L
<j> = 0 on T},
e L2(Q),
V i,j, 1 < i,j <
with 9(0.) = {(p e C^iU), 4> has a compact support in Q}. From these spaces we also define Vg = {y\ye{H\Q)f,
V • v = 0 in Q, v = g on T}.
N
Suppose that f e(H~\Q)) , where H~\Q) = (HU(Q))* = dual space of Hj(Q) and that g e (HV2{T)f, j r g • n dY = 0; it then follows, from Lions, Temam, Ladyshenskaya, Tartar, loc. cit., that (5.8), (5.9) has a unique solution {u, p} e Vg x (L2(Q.)/U). If fe(L2(Q))N and ge(H3'2(T))N, then {u, p} e 2 N (Vg n (H (Q)) ) x (H^nyU) if T is sufficiently smooth. The above u is also the unique solution of the following variational problem (where Vo is obtained by setting g = 0 in the above definition of Vg): Find ueVn such that Vu • Vv dx =
VveK 0 ,
(5.10)
where <•, • > denotes the duality between (if" \Q))N and (H10(Q)f and
I Vu • Vv dx = £ j Vw; • Vu,- dx. 5.3.2.2. ^ new variational formulation of the Stokes problem (5.8), (5.9). For simplicity we suppose that f e(L2(Q))N; we then define Wgs(H\Q))N x Hh(O) by Wg = j {v, 0} G ( H 1 ^ ) ) ' ' x H\(Q\ v| r = g, x=
f V-vwdx, MweH\£l)\. Jn J
(5.11)
248
VII Least-Squares Solution of Nonlinear Problems
We now consider the following variational problem 15 Find {n,\jj} eWg such that v \vu-S\dx
= f f - 0 + S
V{v, cj)}eW0.
(P)
In Glowinski and Pironneau [2] the following has been proved: Theorem 5.1. Problem (P) has a unique solution {u, i//} such that * = 0,
(5.12)
u is the solution of the Stokes problem (5.8), (5.9) (and (5.10)).
(5.13)
EXERCISE
5.1. Prove Theorem 5.1.
Remark 5.1. The potential > introduced above is not at all mysterious. Indeed, the formulation (P) can be interpreted as follows: if v e (H1(Q))N and if F is sufficiently smooth, there exist 0 e H2(Q.) n HQ(Q) and to e (H1(Q))N with V • o) = 0, such that v = -\cj) + co,
(5.14)
and the decomposition (5.14) is unique. In the formulation (P), instead of directly imposing V • v = 0, we try to impose
MT)
V^H-
(5-15)
We then define the following finite element spaces: Hi = {<$>, e C°(Q),
4>U\T
£Pi,VTe
l
(5.16)
H Oh = {cph eHl,(j)h = 0 on T} = Hi n H 0(Q),
(5.17)
Vh = {v, e (C°(Q))2, v, |T e P 2 x P 2 , V T e ^ } ,
(5.18)
i,vft = g f t o n r } 15
.Th}, l
Where Wo is obtained by setting g = 0 in (5.11).
(5.19)
249
5 Numerical Solution of the Navier-Stokes Equations Figure 5.1
(where, in (5.19), gh is a convenient approximation of g whose construction will be discussed in Appendix 3; if g = 0, one takes gh = 0 in (5.19) to obtain VOh),
Wgh = j{v,, 4>h} e Vgh x Hhh,
, dx =
• vhwh dx, V wh e
(5.20) We shall also use the variants of Vgh, Wgh obtained from Vh = {v, e (C°(H))2, v h | r e P , x P : , V T e # , } ,
(5.21)
where # f t is the triangulation obtained from .Th by subdividing each triangle T e 2Th into four subtriangles (by joining the mid-sides; see Figure 5.1). In the above definitions, Pk denotes (as usual) the space of the polynomials in two variables of degree < k. 5.3.3.2. Definition of the approximate problems. Characterization of the approximate solutions. We approximate (P) of Sec. 5.3.2.2 (therefore, the Stokes problem (5.8), (5.9)) by: Find {uh, iph} e Wgh such that
v f Vu, • \vhhdx dx = f f • (v, + Jn
dx,
Jn
V{vh,4>h}eWOh.
(PJ
In Glowinski and Pironneau [2] one may find the proof of the following: Theorem 5.2. Problem (Ph) has a unique solution {uh, ^/h) which is characterized by the existence of a discrete pressure ph e H\ such that
{\Ph-\whdx=
[f-\whdx,
Vw.ef
v f Vu, • Vv, dx = f ( - \ph + f) • v, dx, •In
V
Jn
{uh, i//h} E Wgh (with >j/h*Oin general). EXERCISE
(5.22) VO ,
(5.23)
Oh
(5.24)
5.2. Prove Theorem 5.2.
5.3.3.3. Uniqueness of the discrete pressure. Convergence of the approximate solutions. We introduce the notation 1/2
lo,n
250
VII Least-Squares Solution of Nonlinear Problems
The following lemma (proved in Bercovier and Pironneau [1] and Glowinski and Pironneau [2]) plays a fundamental role in the study of the convergence. Lemma 5.1. Suppose that the angles of 2Th are bounded from below by 90 > 0, independent ofh. Also suppose that V T e ,9~h, T has at most two vertices belonging to P. Then if Hi and VOh are defined by either (5.16) and (5.18), (5.19), or (5.16) and (5.21), (5.19) (with gh = 0 in (5.19)), there exists Co independent ofh such that
V^e4
k*| 1>n
d x
W^»
.
(5.25)
A first consequence of Lemma 5.1 is given by the following: Theorem 5.3. Suppose that (5.25) holds. Then the discrete pressure ph occurring in (5.22)-(5.24) is unique in H},/U (i.e., up to an arbitrary constant). For a proof see Glowinski and Pironneau [2]. EXERCISE
•
5.3. Prove Theorem 5.3.
A fairly complete discussion of the convergence properties of the approximate solutions is done in Glowinski and Pironneau [2]; in the two following theorems we summarize the results which have been obtained there: Theorem 5.4. Suppose that Q is a bounded polygonal convex domain of U2 and that {^~h}h obeys (5.15) and the statement of Lemma 5.1; then the following error estimates hold (i/g has been conveniently approximated by gh), where {ah, [//,,} is the solution of(Ph) and ph is the corresponding discrete pressure: (i) If Hi and Vh are defined by (5.16), (5.18), and ifn e (H3(Q))2, p e H2(Q)/R, we have K
\\Ph ~
- u|| (ffl(n)) 2 < C7i 2 (||u|| (H 3 (n)) 2 + ||p|| H 2 (n)/R ),
^ C/i(||u|| ( H 3 ( n ) ) 2 +
PWHHCWM
(5.26)
( 5 - 27 )
WPWHHO)!®)-
II^IIH'(O) < C/z2(||u||(H3(n))2 + \\p\\HHnm).
(5.28)
l
(ii) If H h and Vh are defined by either (5.16), (5.18) or (5.16), (5.21) and if u e (H 2 (Q)) 2 , p e H\n)/U, we have )/H)-
)
))
)/
(5-29)
(5.30)
In (5.26)-(5.30), C denotes various quantities independent ofu, p, and h. Theorem 5.5. If the hypotheses on Q and {-9~h}h are those of Theorem 5.4, then the following error estimates hold:
5 Numerical Solution of the Navier-Stokes Equations
251
(i) If Hi and Vh are defined by (5.16), (5.18), and ifu e (H3(£l))2, p e H2(O)/U, we have \\Ph - P\\ma)m ^ C/j2(||u||(H3(n))2 + \\p\\H2mu),
(5.31)
3
| | K + Vi/O - u | | ( L W < C7i (||u|| ( H W + ||p||H2(n)/R). (5.32) (ii) If Hi and Vh are defined by either (5.16), (5.18) or (5.16), (5.21) and if u e (H2(Q))2, p e H^ty/U, we have ( 5 - 33 )
\\Ph ~ PWmnyut ^ C7i(||u||(H2(n»2 + \\p\\HHn)l«l ||(uh + ViA,) - u||(L2(S2))2 < C/i2(|ju||(H2(n))2 + \\p\\HHam).
(5.34)
In (5.31)—(5.34), C denotes various quantities independent ofu, p, and h. We observe that the error estimates (5.26)-(5.34) are of optimal order. For more information concerning the convergence, see Glowinski and Pironneau [2], where the extension to finite element approximations of (5.8), (5.9) by rectangular elements is also considered (in Sec. 7). Remark 5.2. Let X be a Banach functional space containing the constant functions; we then define \\v\\xlu,M ve X, by
5.3.4. A mixed finite element approximation of the stationary Navier-Stokes equations We now consider the steady Navier-Stokes equations - vAu + (u • V)u + \p = f in Q,
(5.35)
V • u = 0 in Q,
(5.36)
u = g on T. Using the notation of Sec. 5.3.3, we approximate (5.35)-(5.37) by:
(5.37)
Find {uh, \jjh] e Wgh such that, V {vh,
v f Vu, • Vv, dx + f ( i v V)uft" (VH + W dx= f f • (v, + \(j,h) dx. (5.38)! Jn
Jn
-In
Actually to every solution of (5.38)x we can associate a discrete pressure phsHl, uniquely defined in HJ./R if (5.25) holds, and such that (5.38)! is equivalent to
\\ph-\whdx+ Jn
f (u, • V)uft • VWfcdx = \f-\whdx, Jn Jn
v | V i v Vv, dx + I (uh • V K - v h d x + \ Y P h - v h d x = Jn
Jn
\fyheVOh,
Jn
{uh,iph}eWgh,
pheHl
VwheHhh, f f • vft dx,
Jn
(5.38)2
252
VII Least-Squares Solution of Nonlinear Problems
In Le Tallec [2], [3] it is proved that if some reasonable assumptions on the smoothness of u and p are satisfied, then the estimates (5.26)—(5.31), (5.33) still hold; it is very likely that (5.32), (5.34) also hold, but proving these estimates is still an open problem. 5.3.5. A semidiscrete approximation of the time-dependent Navier-Stokes equations Using suitable finite element bases for the spaces defined in Sec. 5.3.3.1, it is possible to reduce the unsteady Navier-Stokes equations to a system of ordinary differential equations. Suppose that the time-dependent problem under consideration is du — - vAu + (u • V)u + \p = f in Q,
(5.39)
V • u = 0 in Q,
(5.40)
u = g(=g(x,0)onr,
(5.41)
u(x, 0) = u o (x)inQ;
(5.42)
we then approximate (5.39)^(5.42) by: Find {uh(t), \j/h(t)} e Wgh(t) a.e. in t such that
f ^.yhdx
Jn ot
+v (
Jn
f-(v» + S
V{vh,(l>h}£WOh,
(5.43)
uft(0) = uOh e Vh given (uOh: approximation ofu0). In (5.43) the space Wgh(t) is defined as follows: Wgh(t) = | { v h , h • Vwh dx
J
(5.44)
with Vgh(t) = {yheVh,yMr
= gh(t)},
(5.45)
gh(t) being a suitable approximation of g(f). The approximate problem (5.43) is not directly suitable for computation, and for this purpose a convenient time discretization is required in order to obtain a fully discrete approximate problem; several such time discretization schemes are given in Sec. 5.4.
5 Numerical Solution of the Navier-Stokes Equations
253
5.4. Time discretization of the time-dependent Navier-Stokes equations We follow the presentation of B.G. 4P. [1] (see also Le Tallec [2] and Periaux [2]); in the sequel, k = At is the time discretization step. We only consider fully implicit schemes, since the stability condition of the semi-implicit schemes that we have tested (on realistic problems) seems to be a severe limitation with regard to today's computers (a description of such semi-implicit schemes is given in B.G. 4P. [1], Le Tallec [2], and Periaux [2]). The following considerations are, of course, closely related to Chapter III of this book. 5.4.1. An ordinary implicit scheme This scheme is defined as follows (ujj approximates uh(nk)): < = u0/, (see (5.43));
(5.46)
+
then for n > 0, we obtain uJJ * from ujj by solving
f
Jn
K K
~K-yhdx
K
+ v (\u"h+1-\vhdx + f K Jn
Jn
=f
Jn Jn (5.47) m
where (see (5.44), (5.45)) W
m
gh
= Wgh(mk), and f
= f(mk).
+i
To obtain \Th from ujj in (5.47), we have to solve a finite-dimensional nonlinear problem very close to the steady discrete Navier-Stokes problem (5.38). The above scheme has a time truncation error in O(At) and appears to be unconditionally stable. 5.4.2. A Crank-Nicholson implicit scheme This scheme is defined as follows: < = uOh;
(5.48)
then for n > 0, we obtain ul+1from u^ by solving (* "" " Jn
k
~UU*-Yft dx + v ( Vu£+ 1/2 • VvA dx * Jn
+ f(ur Jn
fn
1/2
• (v, + V0 fi ) dx,
V {v,, 4>h} e WOh,
J
{unh+m, M+1/2} £ WgH
1/2
( = Wah((n + l/2)k)),
(5.49)
254
VII Least-Squares Solution of Nonlinear Problems
where <+1)-
(5.50)
Since u£ +1 = unh + 2(uJ|+1/2 - uj), we can eliminate unh+1 in (5.49) and therefore reduce this problem to a variant of the discrete Navier-Stokes problem (5.38). The above scheme has a time truncation error in O(\At\2) and appears to be unconditionally stable. 5.4.3. A two-step implicit scheme The scheme is defined by u£ = uoh, ul given;
(5.51)
then for n > 1, we obtain u£ +1 from uj|, u^" 1 by solving
(5.52) To obtain nj, from u°, we may use either one of the two schemes discussed in Sees. 5.4.1 and 5.4.2 or one of the semi-implicit schemes described in B.G. 4P. [1]; scheme (5.51), (5.52) appears to be unconditionally stable and its truncation error is in O(\ At\2). 5.4.4. Alternating-direction methods Previously we have used alternating-direction methods to solve various kinds of steady nonlinear problems (cf. Chapter IV, Sec. 2.6.6 and also Sec. 3.2.4 of this chapter); actually these methods are also very useful for solving timedependent problems and most particularly the unsteady Navier-Stokes equations as indicated by the two methods described below (and the corresponding numerical experiments). 5.4.4.1. A Peaceman-Rachford alternating-direction method for solving the unsteady Navier-Stokes equations. The method (inspired by Peaceman and Rachford [1]) is denned, with 0 < 9 < 1, by < = uOh;
(5.53)
5 Numerical Solution of the Navier-Stokes Equations
255
then for n > 0, we obtain uj| + 1 / 2 , ujj +1 from u£ by solving I
/l
rt
— *^h
j-j^. Jn
-(- 1 17
I
\hdx + 9v
vujj
{
' • Vvh dx + (1 — 9)v
Jn
K/2 f
+1/9
f,
+
Vu^ • Vvft dx Jn
(uj, • v)uj] • {\h + Vcpj,) dx =
i
Jn
, x ,
' • (v,, + \(ph) dx,
,
..-.
(5.54)j
Jn
which is equivalent to f
+1/9
Jn
vpJJ
f
' • vWfc dx + I (u^ • V)u^ • Vw h d x
Jn
f n+l/2
1
Jn
f
U
^
_
..n
Jn
r
r
+ 0v Jn
VuJ|+1/2 • Vvft dx 4-
Vp|J +1/2 -v,
Jn
f Jn = |F+1/2-v,dx,
VY,6I
and then r u"
Jn
"
— n" + i / z K
v /^-J
*
+ f« Jn
+]
vhdx + (l-6)v
r
Jn
VuJ+1-
• V)u^+1 • \h dx + 6v f \unh+
+ \\pnh+ll2-vhdx= Jn
dx
Ja
Jn
(fn+1-\hdx,
V v e VOh, a f ' e F ; ; ^ F9,((« + l)/c)).
(5.55)
The above scheme (whose practical implementation will be discussed in Sec. 5.4.4.3) has a time truncation error in O(At) and appears to be unconditionally stable in practice. The equivalence between (5.54)t and (5.54)2 is left to the reader as an exercise (see also Sec. 5.7.2.3).
256
VII Least-Squares Solution of Nonlinear Problems
5.4.4.2. A second alternating-direction method for solving the unsteady NavierStokes equations. This is the variant 1 6 of (5.53)—(5.55) denned as follows (again with 0 < 9 < 1): < = u0A; then for n > 0, we obtain
nnh+1/4,
<
+3/4
.n+l/4 _ u «
h
^-— Jn
, nl
(5.56) from u"h by solving
Vur i/4 -
-.vhdx + (l-0)v
K/4
+l
Jn
+ 9v f Vu"h • \yh dx + f « • V)u^ • (y
(= ^ ( ( » (5 57)1
-
which is equivalent to
f \p"h+1/4 • Vw, dx + f (u; • V)u^ • Vwft rfx Jn Jn Jn
Jn
+ (1 - fj)v f Vu£ + 1 / 4 • Vv,rfx+ f VpJ Jn
Jn
+ dv \ \u"h • \vh dx + f (uj • V)u^ • vft rf Jn Jn
= (V1/4-Mx,
vv,6yOi,
Jn then ..n+3/4 Ufi
—
,,n+l/4 lit
6v
Jn
Vun+3 / 4 -
n f
Jn
\yhdx
Jn n+l/4
f n+3/4
Jn (5.58)
' Inspired to us by Leveque and Oliger [1] (see also Leveque [1]).
5 Numerical Solution of the Navier-Stokes Equations
257
and finally /*
I
II 0 "*" 1
h
i
n n +3/4
/•
— **h i /^
i n
/
r\
v
/
+ 9v (\u"h+3'4-\\hdx+
I i
+1 n
n
f (u2 +3/4 -VK +3/4 '
-L' (5.59) Scheme (5.56)-(5.59) has a time truncation error in O(\At\2) and appears to be unconditionally stable in practice; the practical implementation of (5.56)(5.59) is discussed in Sec. 5.4.4.3. 5.4.4.3. Practical implementation of (5.53)-(5.55) and (5.56)-(5.59): further remarks. The main advantage of the alternating-direction methods (5.53)(5.55) and (5.56)-(5.59), compared to the methods of Sees. 5.4.1, 5.4.2, and 5.4.3, is that they decouple the two main difficulties of the Navier-Stokes equations, which are nonlinearity and incompressibility. In both methods, the nonlinear problems solved at each time step (namely (5.55) and (5.58), respectively) are in fact finite-dimensional variants of the nonlinear problem (3.56) (cf. Sec. 3.4, Exercise 3.1), and therefore they can be solved by least-squares conjugate gradient methods very close to those methods used in Sec. 3.4 for solving the model problem (3.1). On the other hand, problems (5.54) and (5.57), (5.59) are closely related to the discrete Stokes problem (Ph) (cf. Sec. 5.3.3.2) whose numerical solution will be discussed in Sec. 5.7. Remark 5.3. With respect to storage requirements, the optimal values 17 of 9 are 9 = £ for (5.53)-(5.55) and 6 = i for (5.56)-(5.59); these choices will be justified in Sec. 5.7 and are directly related to the properties of the methods used to solve the nonlinear problems (5.55), (5.58) and the discrete Stokes problems (5.54), (5.57), and (5.59). Remark 5.4. From our computational experiments, (5.55) (resp., (5.58)) is the most costly 18 step (in term of computational time) of method (5.53)-(5.55) (resp., (5.56)-(5.59)); this observation justifies the choice made in (5.56)-(5.59) for the symmetrization process leading to an O(\ At|2) time truncation error; indeed the cost of step (5.57)-(5.59) is practically the same as the cost of step (5.54), (5.55).
17 18
Actually these values also lead to shorter computational times. About 5 to 10 times.
258
VII Least-Squares Solution of Nonlinear Problems
Remark 5.5. The alternating-direction method (5.53)-(5.55) is in fact a discrete variant of the following semidiscrete scheme for integrating the time-dependent Navier-Stokes equations (5.39)-(5.42): u° = u0; then for n > 0, we obtain {u" +1/2 , pn+1'2}, \pn+
k/2
m
(5.60)
an+1 from u" by solving
- (1 - 0)vAu" + (u" • V)u" = f"+m
in Q,
V • u" + 1 / 2 = 0 in Q, u" + 1 / 2 = g((n + l/2)fc) on T, (5.61) and 1
(i _ 0)vAu"+1 + (u"+1 • V)u"+1 - 0vAu" +1/2 + Vp" +1/2
IL"
= f"+ x in Q, u" + ' = g((n + l)fe) on T. (5.62)
Similarly, method (5.56)-(5.59) is the discrete variant of a semidiscrete scheme that the reader may formulate by himself. Remark 5.6. Alternating-direction methods are closely related to fractional step methods (basic references for fractional step methods are Yanenko [1] and Marchouk [1]). Actually alternating-direction and fractional step methods have been extensively used for quite a long time for solving time-dependent partial differential equation problems; concentrating particularly on NavierStokes equations for incompressible viscous fluids, we shall mention Chorin [l]-[4], Fortin [3], and Temam [1] (see also the references therein). Remark 5.7. A fractional step variant of the following alternating-direction method has been used by Benque, Ibler, Keramsi, and Labadie [I] 1 9 to solve the unsteady Navier-Stokes equations; the alternating-direction method is defined by 20 : u° = u0: then for n > 0, we obtain u" +1/2 , p"+1/2 n+l/2 _
— k/2
and u" + 1 from u" by solving
n
vAu" +1/2 + Vp" +1 ' 2 + (u" • V)u" = f"+1/2 in Q, V • un+ m = 0 in Q, u" + 1 / 2 = g on T,
19 20
(5.63)
See also P i r o n n e a u [ 1 ] . We suppose, for simplicity, that g ( = u | r ) is independent of t.
(5.64)
5 Numerical Solution of the Navier-Stokes Equations
259
and then, — k/2
h (u" +1/2 - V)u"+1 - vAu" +1/2 + Vp" +1/2
= F + 1 in fi, u"+1 = gonT_, (5.65) where
F_ = {x|xer,g(x)-n(x)<0} (n(x) is the unit vector of the outward normal at T, at x). Problem (5.65) is a linear first-order system which has been solved in Benque, Ibler, Keramsi, and Labadie [1] by a method of characteristics; on the other hand, problem (5.64), which is quite close to the Stokes problem (5.8), (5.9), has been solved in the above reference by the methods of Sec. 5.7 via the finite element approximations of Sec. 5.3. 5.5. Least-squares conjugate gradient solution of the stationary Navier-Stokes equations: (I) The continuous case 5.5.1. Generalities: synopsis
The various discrete Navier-Stokes problems that we have encountered in Sees. 5.3.4,5.4.1,5.4.2, and 5.4.3 are in fact finite-dimensional variants (obtained by finite element approximations) of the following nonlinear problem (with a > 0) au - vAu + (u • V)u + \p = f in Q, V • u = 0 in fi, u = g onT (with \gndT
= 0), (5.66^
which is clearly a generalization of the steady Navier-Stokes problem (5.35)(5.37). Problem (5.66)l can also be formulated as a nonlinear variational equation by: Find ueVg such that r r r r a
u • v dx +
Vu • Vv dx +
v • (u • V)u dx =
f • v dx,
VVGF0
(5.66)2 (see Sec. 5.3.2.1 for the definition of Vo, Vg). Following B.G. 4P. [1], Bristeau, Glowinski, Periaux, Perrier, and Pironneau [1], Bristeau, Glowinski, Mantel, Periaux, Perrier, and Pironneau [1], and Periaux [2], we shall describe, in Sec. 5.5.2, a least-squares formulation of
260
VII Least-Squares Solution of Nonlinear Problems
(5.66) and in Sec. 5.5.3 a conjugate gradient algorithm for solving the leastsquares problem. The discrete variants of the methods described in Sees. 5.5.2 and 5.5.3 will be described in Sec. 5.6. 5.5.2. A least-squares formulation of (5.66) A natural least squares formulation of (5.66) is Find u € Vg such that
J(u) < J(\),
(5.67)
with
= ? f I4WI2 dx + - [ | V$(v)|2 dx,
(5.68)
where £(v) (==£) is a function of v through the state equation <x£ - vA^ + V?r = <xv - vAv + (v • V)v - f in Q, V • \ = 0 in Q, £, = 0 on V. (5.69) We observe that t, is obtained from v through the solution of a Stokes problem, and also that the formulation (5.67)-(5.69) is a natural generalization of (3.37) in Sec. 3.3.2. The above least-squares formulation is justified by the following obvious proposition. Proposition 5.1. Suppose that {a, p} e Vg x (L2(Q)/R) is a solution of (5.66); then u is also a solution of the least-squares problem (5.67), and we have, for the corresponding pair {!;, n}, % = 0,
n = -p;
(5.70)
we also have J(u) = 0. 5.5.3. Conjugate gradient solution of the least-squares problem (5.67)-(5.69) 5.5.3.1. Description of the conjugate gradient algorithm. A conjugate gradient algorithm for solving the least-squares problem (5.67)-(5.69) is defined as follows. Step 0: Initialization a°eVg given;
(5.71)
then compute z° as the solution of the linear variational equation a (z°-x\dx
+ v f Vz° • Vri rfx = <J'(u°), ri>,
VrieF0,
zoeVo, (5.72)
5 Numerical Solution of the Navier-Stokes Equations
261
and set w° = z°.
(5.73)
For m > 0, assuming that um, zm, wm are known, compute u m + 1 , z m + 1 , w m + 1 fry the following. Step 1: Descent Compute Xm = Arg min J(\f - Awm),
(5.74) (5.75)
m+1
Step 2: Construction of the new descent direction. Define z zm+1eV0,
a [zm+1-xydx Jn
as the solution of
(\zm+1-\i\dx
+ v Jn
= <J'(u ra+1 ),ti>,
VrieK 0 ;
(5.76)
then (Polak-Ribiere strategy) • V(zm+ * - zm) dx a$n\zm\2dx
+ v$n\\z\dx
'
(
'
and finally w m + 1 = z m + 1 + ym+1w'",
(5.78)
m = m + 1 go to (5.74). Remark 5.8. To obtain z m + 1 from u m + 1 , via (5.76), we also have to solve a Stokes problem (written in a variational form). 5.5.3.2. Calculation ofJ'(am+ J ) and zm+1. A most important step in order to use algorithm (5.71)-(5.78) is the calculation of J'(u m+ x ); due to the importance of this step, we shall show this calculation in detail. We have
SJ = '(v), S\} = a f £ • b\dx + v f V£ • \3Z,dx;
(5.79)
to express <5^ as a function of <5v, we observe that (5.69) also has the following variational formulation: \ e Vo and V t| e Vo, we have a
Jn
\ • \\ dx + v
Jn
V£ • \r\ dx = a \ y-r\dx Jn
+ v
Jn
\v\r\dx
+ f ii • (v • V)v dx - f f • ti dx, Jn Jn
(5.80)
262
VII Least-Squares Solution of Nonlinear Problems
which in turn implies: 5% e Vo and V r\ e Vo, we have <x f b\ • i[ dx + v f V<5£ • Vt] dx = a t 5\-i\dx Jn Jn Jn
+ v I* \8\ • Vii dx Jn
+ f ti • (<5v • V)v dx + f TI • (v • \)3\ dx. Jn Jn
(5.81)
Since \ e Vo, we can take r\ = i; in (5.81); it then follows, from (5.79), that dJ = (J'{y), d\) = a f Z, • d\ dx + v f V^ • \d\ dx Jci Ja + f § • (5v • V)v dx + f % • (v • V)c5v
rfx.
(5.82)
Finally, from (5.82) it follows that J'(v) can be identified with the linear functional on Vo denned by ti -> a f \ • r\ dx + v f V^ • \r\ dx + f % • (n • V)v dx + f % • (v • \)r\ Ja Jn JQ Jn
dx. (5.83)
m+ i
m+1
From (5.83) it then follows that to compute z from u , we have to first solve the Stokes problem (5.69), (5.80) with v = u m + 1 , which gives ^ m+ J ; then from (5.83) we have V x\ e Vo
<J'(«m+1),ri> = a [%n+1-r\dx + v fv^ m + 1 -Vtidx Jn Jn + f ^ m+ x • (ti • V)u m+ ' dx+ Jn
f ^ m + i • (u m+x • V)i] dx. Jn
(5.84)
Finally we obtain z m + 1 from (5.76), (5.84). To summarize, at each iteration of the conjugate gradient algorithm (5.71)(5.78), we have to solve several Stokes problems, namely: (i) the Stokes problem (5.69), (5.80) with v = u m + 1 to obtain £ m + 1 from um+l.
(ii) the Stokes problem (5.76) to obtain (via (5.84)) z m + 1 from u m+ \ ^ m+ x; (iii) the solution of the one-dimensional problem (5.74) requires several calculations of J, i.e., several solutions of the state equation (5.69), (5.80) (which is a Stokes problem). Thus an efficient "Stokes solver" will be an important tool for solving the Navier-Stokes equations by algorithm (5.71)—(5.78) via the least-squares formulation (5.67)-(5.69). The description of such solvers is given in Sec. 5.7.
5 Numerical Solution of the Navier-Stokes Equations
263
5.6. Least-squares conjugate gradient solution of the stationary Navier-Stokes equations: (II) The discrete case In this section we consider the discrete analogue of Sec. 5.5. We apologize in advance for the ponderous notation to be used (unfortunately this complicated notation seems to be inherent to mixed finite element methods). 5.6.1. Formulation of the basic problem The discrete Navier-Stokes problems encountered in Sees. 5.3.4, 5.4.1, 5.4.2, and 5.4.3 have the general formulation: Find {uh, il/h} e Wgh such that cc I uh • yh dx + v I VuA • \vh dx + f (u, • V)uh • (v, + \
+ [flh-(vh Jn
+ Y4>h)dx,
V K , >„} e Wo,,
(5.85)
where a. > 0, v > 0. 5.6.2. A least-squares formulation o/(5.85) By analogy with Sec. 5.5.2, we shall consider the following least-squares formulation: Min
Jh{yh,(j>h)
(5.86)
{vh,
with
UVH, <W = ^ f IS*I2 dx + "- f |V^| 2 dx, l
^ Jn
(5.87)
Jn
where \h is a function of {\h, <j)h} according to the state equation
a Jn +
Jn
\h • i\hdx + v
Jn
\$h • \r\h dx = a
(v* • V)vft • (% + \wh) dx -
V { % , mh} e WOh.
Jn
vh-r\hdx
\ fOh-r\hdxJn
Jn
+ v Jn
V\h • \r\h dx
fu • (x\h + V o J dx, (5.88)
264
VII Least-Squares Solution of Nonlinear Problems
The least-squares formulation (5.86)—(5.88) is justified by the following: Proposition 5.2. Suppose that {uh, \j/h} e Wgh is a solution o/ (5.85); {uh, \jjh) is also a solution of the least-squares problem (5.86), and we have, for the corresponding pair {%h, xh},
we also have Jh(nh, i/^) = 0. 5.6.3. Computation of the gradient J'h From (5.87), (5.88) we have dJh = a\$h- 5$, dx + v f nh • S5%h dx, Jn Jn where {5£,h, 5xh} is the solution of:
(5.89)
{&$h, o~Xh} e WOh and V {i\h, coh} e WOh, we have
a \d%h-r\hdx Jn
+ v
Jn
+ \(foh-\)vh-(rh Jn
V<5^• V% dx = cz 5\h • r\hdx + v \S\h Jn Jn + Vcoh)dx+
((vh-V)dvh-(r\h Jn
dx
+ Vo)h)dx.
(5.90)
Since {%h, xh} e WOh, we have, from (5.90) (taking {%, coh} = {£,„, xJ),
a Uh-S2,hdx + v fv^-V^dx = a \^-Syhdx + v f \$h-\5yhdx Jn
Jn
Jn
Jn
+ f ( H • \)yh • ( ^ + VZft) dx + f (v, • \)5vh • fe + \Xh) dx.
(5.91)
From (5.91) it then follows that J'h(\h, (j)h) may be identified with the linear mapping from WOh to U defined by <^(v*. 4>h), {%,<»»}> = « f ^ • % dx + v f V^fe • Vt|fc dx Jn
Jn
f
dx,
Jn (5.92) 5.6.4. Conjugate gradient solution of the least-squares problem (5.86)^(5.88) In this section we shall describe a conjugate gradient solution of the leastsquares problem (5.86)-(5.88). The algorithm to be considered is in fact a discrete analogue of algorithm (5.71)-(5.78) in Sec. 5.5.3, given by the following.
5 Numerical Solution of the Navier-Stokes Equations
265
Step 0'. Initialization {nl,^h}eWgh
given:
(5.93)
then compute {z°, 9°} as the solution of the variational
problem
a f x°h •% dx + v f Vz° • V % dx = <J'h(u°h, «K°), {r\h,
(5.95) 1
For m > 0, assuming { < , i/^H e W^^, {z|*, 0^}, {w^ , T^} E WOh known, compute W + S ^ + 1 } , {zr+ 1 , ^ + 1 } , {w^ +1 , T ^ + 1 } by the following. Step 1: Descent Xm = Arg min JM
~ A<, ^
u m+l
=
K+1
= W-
- ATJR,
A™T?.
(5.96)
(5.98)
Step 2: Construction of the new descent direction. Define {zjj1"1"1, 0™+1} as the solution of the variational problem
{z^+1,0^+1}€^OA,
V {iu, a>fc} e Wo*.
(5.99)
w/zere J^(M™ + S IA™+ X ) is obtained from (5.92), and
_ « In z r 1 • W + 1 ~ *X) dx + v Jn Vz?+
a j a W r ^ + vjnivzjri2^ w^ + 1 = z™+1 + ym+1v%, +1
T™
+1
= 0™ + y m+1 T^,
(5.101) (5.102)
m = m 4- 1 go to (5.96). It is clear that each iteration of algorithm (5.93)-(5.102) requires the solution of several discrete Stokes problems: (i) one to obtain {££+1, xt+i}
v/, = <
+1
,
+ 1
from {u™+*, \//%+1}; in fact this is (5.88) with
;
(ii) problem(5.99)toobtain{z^ +1 , ^ + 1 } f r o m { u ^ + 1 , ^ + 1 } , { ^ + 1 , ^ + 1 } ; (iii) the solution of the one-dimensional problem (5.96) requires several calculations of Jh, i.e., several solutions of the state equation (5.88) (which is a discrete Stokes problem). The solution of these discrete Stokes problems will be discussed in Sec. 5.7.
266
VII Least-Squares Solution of Nonlinear Problems
5.7. On efficient solvers for the Stokes problem In this section we shall discuss the solution of the various Stokes problems encountered in the previous sections, by either direct or iterative methods. 5.7.1. The continuous case We follow B.G. 4P. [1]. 5.7.1.1. Synopsis. From Sees. 5.4 and 5.5 it follows that efficient solvers for the Stokes problem au - vAu + Vp = f infi,
V u = 0inQ,
u = g on F
(5.103)
will be advantageous for the numerical solution of the steady and unsteady Navier-Stokes equations. In (5.103), a = 0 for the steady case and a > 0 for the unsteady case. We shall describe a method for solving (5.103). This method reduces the solution of the Stokes problem (5.103) to the solution of a finite number of Dirichlet problems and to the solution of a boundary integral equation. The finite element implementation of this method will be discussed in Sec. 5.7.2 (see also Glowinski and Pironneau [3], Glowinski, Periaux, and Pironneau [1], and B.G. 4P. [1, Sec. 8]). The reader mainly interested in applications (or frightened off by too much functional analysis!) can skip the rest of Sec. 5.7.1 and move on to Sec. 5.7.2. 5.7.1.2. Principle of the method. Let us define
^ 1/2 (F) = L H1I2(T), I For the definition and properties of the Sobolev spaces HS(T), seU, see Lions and Magenes [1] and Necas [1]. The decomposition properties of the Stokes problem follow directly from: Theorem 5.6. Let XeH" 1/2 (F) and let the operator A: H~ 1/2(F) -> H 1/2 (F) be defined implicitly by the following cascade of Dirichlet problems: Apx = 0 in Q,
px = X on F,
auA — vAuA = —'Spi in Q, -Mix = SwxinQ.,
uA = 0 on F, ^ = 0onF,
(5.104) (5.105) (5.106)
and then AX= Then A is an isomorphismfrom H form a(-, •) defined by
-
dip. dn
(5.107)
1/2
(F)/IR onto #f 1/2 (F). Moreover, the bilinear
a{X, ii) = (AX, /i>,
Vn e H
1/2
(F)
(5.108)
5 Numerical Solution of the Navier-Stokes Equations
267
(where <(•,•) denotes the duality pairing between H1I2(T) and H~ 1/2 (F)) is continuous, symmetric, and H~ ll2(r)/U-elliptic. We recall that a bilinear form denned over V x V, where V is a Hilbert space, is F-elliptic if there exists /? > 0 such that a(v, v) > p\\v\\2, V v e V. For the proof of Theorem 5.6, see Glowinski, Periaux, and Pironneau [1] (a variant of it concerning the biharmonic problem is avaiable in Glowinski and Pironneau [1]). EXERCISE
5.4. Prove Theorem 5.6.
Application of Theorem 5.6 to the solution of the Stokes problem (5.103). Let us define p0, u 0 , \j/0 as the solutions of, respectively, Ap0 = V • f in Q,
po = 0 on F,
au0 — vAu0 = f — \p0 in Q, -A\j/0 = V u o i n Q ,
u 0 = g on F, i^0 = 0 on T.
(5.109) (5.110) (5.111)
The following result is established in Glowinski, Periaux, and Pironneau [1]: Theorem 5.7. Let {a, p} be a solution of the Stokes problem (5.103). The trace X = p\r is the unique solution of the linear variational equation (E)
XeH-lt\T)IU,
EXERCISE
{AX, fi> = / ^ , A
V n e H" 1 / 2 (r)/R.
(5.112)
5.5. Prove Theorem 5.7.
From the above theorem it follows that the solution of the Stokes problem (5.103) has been reduced to 2N + 3 Dirichlet problems 21 (N + 2 to obtain \j/0, N + 1 to obtain {u, p} once X is known) and to the solution of (E) which is a kind of boundary integral equation. The main difficulty in this approach is the fact that the (pseudo-differential) operator A is not known explicitely; actually the mixed finite element approximation of Sec. 5.3 has been precisely introduced (in Glowinski and Pironneau [4], [5]) to overcome this difficulty, extending to the Stokes problem an idea discussed in Glowinski and Pironneau [1] for the biharmonic problem. 5.7.2. The discrete case 5.7.2.1. Introduction: formulation of the basic problem. From Sees. 5.4 and 5.6 it follows that efficient solvers for the discrete Stokes problem:
21
N = 2 i f H c U2;N = 3 i f Q c R 3 .
268
VII Least-Squares Solution of Nonlinear Problems
Find {uh, \j/h} e Wgh such that, V {\h,
El = Hhh © Mh; we set Nh = dim(Mfc). Various Mh may be used, but in the following we shall only consider Mh defined by Hi = HlOh 0 Mh; then, V fih e Mh, we have HH\T = 0, V T e ^ such that dT n 3Q = 0.
(5.114)
The space Mh is uniquely defined by (5.114), and we observe that if (ih e Mh, then /xh vanishes outside a neighborhood of 8Q, whose measure ->0 with h. Therefore Mh can " almost" be considered to be a boundary space, and in fact it will play the role played by H~il2(F) in Sec. 5.7.1. We also observe that \ih&Mh is completely determined by its trace on <3Q, and in fact by the values it takes at the boundary nodes. 5.7.2.3. Equivalence o/(5.113) with a variationai problem in Mh. Let XheMh; from Xh we define ph, uh, \j/h as the solutions of the cascade of discrete variationai Dirichlet problems: Find ph e Hi such that ph — Xh e Hlh and
[\ph-\qhdx
= O,
VqheHhh.
(5.115)
Jn Find uh e VOh such that a I uh • \hdx + v | Vuh • \\h dx = - \ \ph • \h dx, Jn Jn Jn
Vv ft e Voh. (5.116)
Find \jjh e HQH such that
f Jn
r Jn
x
5 Numerical Solution of the Navier-Stokes Equations
269
Then from (5.115)-(5.117) we define ah: Mh x Mh -> U by ah{Xh, fih) = — (\\j/h + nh)-\nhdx. (5.118) Jn The following analogue of Theorem 5.6 of Sec. 5.7.1.2 is proved in Glowinski, Periaux, and Pironneau [1]: Theorem 5.8. Assume that V T' e ^ , T' has at most one edge supported by 8Q.;
(5.119)
then ah(-, •) is bilinear, symmetric, and positive definite over (MJUh) x (Mh/Uh), where DFSef = {/ih e Mh, fih = constant on 80.}. EXERCISE
5.6. Prove Theorem 5.8.
Referring back to Sec. 5.7.2.1, it is easily proved (see Glowinski, Periaux, and Pironneau [1]) that (5.113) has a unique solution which is also the unique solution of the minimization problem
Min
^ f K | 2 dx + V- f |VvJ 2 dx - f fOft • vA dx
l
Jn
Jn
- J f i * ' ( v * + ^H)dx\.
(5.120)
Since
Wgh = \{yh, 4>h} e Vgh x Hhh, J V0fc• \qh dx = J V• vhqhdxV
qheHl\
(5.120) is a linearly constrained minimization problem (the number of linearly independent constraints is dim(Hj,)). We can therefore associate with (5.120) a Lagrangian functional (see, e.g., G.L.T. [1, Chapter 2], [3, Chapter 2]) &k\
Vh x Hi
xHl^U
defined by
(\tl>k-\qhdxJn
(\-vhqhdx, Jn
(5.121)
where
4>H) =1 ? f IvJ 2 dx + I f | VvJ 2 dx - f fOh • yh dx Jn
*• Ja
Jn
- fflh-(vh Jn
+ \<j>h)dx.
(5.122)
270
VII Least-Squares Solution of Nonlinear Problems
Since (5.120) is a finite-dimensional linearly constrained minimization problem whose solution exists, there is a Lagrange multiplier related to (5.120)-(5.122) (cf., e.g., Rockafellar [1] and also Chapter I, Sec. 7.4.2), i.e., there exists ph e H\ such that {ah,il/h, ph} is a saddle point of S£h over Vgh x H^h x Hi (where {uh, i//h} is the solution of (5.113), (5.120)). Since <£h is extremal at {uh, ij/h, ph}, we obtain (5.123) a \ uh • vhdx + v Vuh • V\h dx + \ \ph • \h dx Jn Jn Jn fo* + fi*)-v»dx,Vv h 6 7Oh, uheVgh, = [s/.xxhqhdx, Jn
VqheHlh.
(5.124) (5.125)
Relations (5.123)-(5.125) characterize {uh, i//h} as the solution of (5.113), (5.120). An important conclusion is that the multiplier ph appears from (5.124) as the discrete pressure. In fact it is proved in Glowinski and Pironneau [2] that if (5.119) holds, then ph is uniquely defined in Hj,/U. The key result concerning Sec. 5.7.2 is the following theorem which links the data gh ( = uh\r), fOh, flh and the trace of the discrete pressure ph: Theorem 5.9. Let ph be the above discrete pressure and Xh the component ofph in Mh. Then if (5,.119) holds, Xh is the unique solution in MJU lh of the linear variational problem (discrete analogue of(E) in Sec. 5.7 .1.2): (EA) Xh e Mh/Uh, ah(Xh, /v> = f (V^o* + "OH) •' i7^^ dx, Jn where
eMJR (5. 126)
pOh,uOh, \j/oh are, respectively, solutions of
hpoH- \qh dx = flh • \qh dx, Jn Jn a \ uOh- \h dx Jn
f.-, i
I
Jn
„ Oh
V qh e HL,
p
(5. 127)
,
hL*"^/
= f(fo* + fi*-Vpo*)-v*
,,
(Vu o ,,<Mx, Jn
V0,6Hj f t ,
uoheVgh, (5.128) ilfOheHbh.
(5.129)
Theorem 5.9 (whose proof is given in Glowinski, Periaux, and Pironneau [1]) is in fact a variant of Theorem 3.2 in Glowinski and Pironneau [1] concerning approximate solution of the biharmonic problem.
5 Numerical Solution of the Navier-Stokes Equations
271
The solution of (Eh) by direct and iterative methods is discussed in Sees. 5.7.2.4 and 5.7.2.5, respectively. 5.7.2.4. Solution of(Eh) by a direct method. This section follows B.G. 4P. [1], Glowinski and Pironneau [3] ,and Glowinski, Periaux, and Pironneau [1]. A. Construction of a linear system equivalent to (Eh) General: As before, Mh is defined by (5.114); let
be a basis of Mh. Then, V \xh eM h ,
Hh=
t/,iWi,
(5.130)
and from now on we shall write Wh = (A*!, . . . , nNh} € UN\
(5.131)
In practice Bh is defined by Bh = { w j ? ^
(5.132)
and (see Fig. 5.2) V i = 1 , . . . , Nh,
wt e Mh,
Wi{Pt)
= 1,
Wi(Q)
=0
V Q vertex of 2Th,
Q # Pt, (5.133)
where we have implicitly assumed (though in practice this is not strictly necessary) that the boundary nodes are numbered first. With this choice for Bh, we have ^ ; = nh{P,) in (5.130), (5.131). Then (EA) is equivalent to the linear system
t ah(Wj, j= i
W j )A ;
= f W o , + u oft ) • Vw ; dx,
l < i < Nh. (5.134)
Ja
Let aij = ah(wj, w,), Ah = ( a ^ < ; , ; < „ „ , bt = | a W M + uoh) • Vw; dx, bh = {b,-}f=i- The matrix Ah is full and symmetric, positive semidefinite. If (5.119) holds, then 0 is a single eigenvalue of Ah; furthermore, if Bh is defined by (5.132), (5.133), then Ker(A») = {y |y e R»», y1=y2
Figure 5.2. The support of wt is shown.
= ... = yNh}.
(5.135)
272
VII Least-Squares Solution of Nonlinear Problems
As to the conditioning of Ah restricted to R(Ah) ( = UNh — Ker(AA)), it can be shown that the condition number v{Ah) is of order h~2 if (5.119) holds. In fact, by analogy with Glowinski and Pironneau [1, Sec. 4] it is reasonable to conjecture that v(A^) = 0(h~J), but we have not been able to obtain this estimate. Construction ofbh. To compute the right-hand side of (5.134), it is necessary to solve the four (five if Q c R3) approximate Dirichlet problems (5.127)(5.129). Due to the choice (5.114) for Mh, the integrals in the definition of the right-hand side of (5.134) involve functions whose supports are in the neighborhood of dQ. only (see Fig. 5.2). Construction of Ah. Ah is constructed column by column according to the relation atj = ah(wj, wt). To compute thejth column of Ah, we solve (5.115)— (5.117) with Xh = w ; and compute a y from (5.118). Thus four discrete Dirichlet problems must be solved for each column (five in R3). Since the matrix Ah is symmetric, we may restrict i to be greater or equal to j . Incidentally, the above observation regarding the choice of Mh still holds, so that the integrals occurring in the aVj are not very costly to compute. B. Solution of(Eh) by the Cholesky method. Assume that (5.119), (5.133) hold. Then it may be shown from (5.135) that the submatrix Ah = (fly)i < ij
(5.136)
(where fhXh = {A 1 ; ..., XNh_ J , b,, = {bu ..., bNh_ y}) by the Cholesky method via a factorization A, = L,U(or Ah = L.DfcU),
(5.137)
where hh is lower triangular and nonsingular (and Dh is diagonal). Let us review the subproblems which arise in the computation of {u,,, ph} via (Eh) if the Cholesky method is used: (i) the four (five if Q a U3) approximate Dirichlet problems (5.127)(5.129): (ii) 4(Nh — 1) approximate Dirichlet problems to construct Ah (5(Nh — 1) if Q c U3); (iii) two triangular systems to compute Xh, namely:
Ufh = K,
U'/A = Vh,
(iv) three approximate Dirichlet problems to obtain ph and ah from Xh (four if Q <= R3). In practice the matrices of the approximate Dirichlet problems should be factorized once and for all. There are two symmetric positive-definite matrices, one to approximate — A using affine elements, one to approximate al — vA
5 Numerical Solution of the Navier-Stokes Equations
273
using quadratic elements (or affine on # A if (5.21) is used); an alternative is to use very efficient iterative solvers to solve these approximate Dirichlet problems, like those described in Glowinski, Periaux, and Pironneau [2], Glowinski, Mantel, Periaux, Pironneau, and Poirier [1], and Glowinski, Mantel, Periaux, Perrier, and Pironneau [1], for example. Remark 5.9. We complete Remark 5.3 of Sec. 5.4.4.3. If one takes 0 = ^ (resp., 0 = x) in the alternating-direction method (5.53)-(5.55) (resp., (5.56)(5.59)), it follows from the decomposition properties of the Stokes problem discussed in Sees. 5.7.1 and 5.7.2 that the Dirichlet operators 22 21 jk — (v/2)A (resp, 2121/k - (v/3)A]) are involved in the Stokes step(s) (5.54) (resp., (5.57), (5.59)). Since the nonlinear step (5.55) (resp, (5.58)) involves the operator 22 21/k — (v/2)A(resp, 21/k — (v/3)A) (if, for example, a least-squares conjugate gradient method 23 is used), we see that with the above values of 9, the same solvers can be used for the Stokes steps and the nonlinear steps of methods (5.53)-(5.55) and (5.56)-(5.59); with other values of 8, this would no longer be true and about twice as much computer storage would be required. 5.7.2.5 Solution of (Eh) by a conjugate gradient method: further remarks. We follow Glowinski and Pironneau [3], [5] and Glowinski, Periaux, and Pironneau [1]. In Sec. 5.7.2.4 we have described a direct method for solving (Eh) (in fact the equivalent linear system (5.134)); actually a possible alternative is to solve (Eh) (and therefore (5.113)) by a conjugate gradient algorithm; this possibility follows from the fact that the bilinear form ah(-, •) in (Eft) (cf. (5.126)) is symmetric and positive semidefmite (from Theorem 5.8) (or equivalently, the matrix Ah in the linear system (5.134) is symmetric and positive semidefinite). Using the notation of Sec. 5.7.2.4, a possible conjugate gradient algorithm is denned by the following. Step 0: Initialization A° e Mh arbitrarily given;
(5.138)
g? = A,r,A 0 - bA,
(5.139)
< = g?.
(5-140)
For n > 0, assuming that Xnh,%"h, z"h are known, we compute X"h+1, g£ + 1 , z £ + 1 by the following.
22 23
In fact their discrete analogous. Like those described in Sees. 3.3, 3.4.
274
VII Least-Squares Solution of Nonlinear Problems
Step 1: Descent
llgiilU2 (5-142) (5-143) Step 2: Construction of the new descent direction (5.144) zj[+1 = 8S +1 +
(5-145)
7« + IZJJ,
n = n + 1 go to (5.141). In (5.138)—(5.145), (•, -)h and ||-||A stand for the standard Euclidian scalar product of UNh and the corresponding norm, respectively (but one could use other scalar products and norms). With the matrix Ah symmetric and positive semidefinite, one can show that the sequence {X"h}n>0 converges 24 to Xh, solution of (Eh); the component of Xh in Uh is that of X°. Implementing (5.138)-(5.145) requires the solution of four Dirichlet problems at each iteration (five if Q <= R3) to compute AfczJJ using the relation
and also the definition of ah(-, •) (given by (5.115)-(5.118)); let us show the details the computation of AhzJ. With z"h = {z?}?^! known (from (5.140) or (5.145)), we first define z\ e Mh by zh - rh zh
^ _ LW)'
(5.i4 6 )
and then from Th we compute n\, v>"h, xl by
nleHl
nl-znheHlh,
J \nnh-\qhdx
= 0,
VqheHbk,
o>nheV0h, a f (o"h-yhdx + v \ \e>nh-\\hdx = - f \n"h-vhdx
(5.147)
Vv ft eF M , (5.148)
Xl • V(j>h dx = I V • &l4>h dx, Jo
1
V
(5.149)
In at most Nh — I iterations (assuming that roundoff errors are not taken into account).
5 Numerical Solution of the Navier-Stokes Equations
275
Then from (5.118) we have (AfcZjJ, thnh)h = ah(z"h, fih)= -
f (Vtf + O • \tih dx
(= £ ( (5.150)
which implies (A,X, z"h)h = ah(z«h, zj) = -
f (VzJ! + coj) • Vzl dx.
(5.151)
Here, also, one should factorize the matrices of the approximate Dirichlet problems. Remark 5.10. The conjugate gradient algorithm (5.138)—(5.145) used to solve the linear problem (Eh) is the classical one (see Hestenes [2] for a detailed analysis of conjugate gradient methods applied to the solution of linear problems); the scalar pn occurring in (5.141) is the solution of the single-variable minimization problem Pne
R,
Ih(rhXnh - p X ) <
hM
where Vy ft e MN\ h(yh)
= M A ^ , yh)h - (bh, yh)h.
We recall that (A,X, zi), = 0,
(gi, «0* = 0,
\/i,j,
VU,
i*j,
i>j;
( 5 - 152 )
the various formulas in (5.141), (5.144) then follow from (5.152). Remark 5.11. Our numerical experiments show that the conjugate gradient algorithm (5.138)-(5.145) converges quickly^even on domains with corners. Typically, convergence is obtained in O(y/Nh) iterations. Remark 5.12. The direct and iterative methods described in Sees. 5.7.2.4 and 5.7.2.5, respectively, have been applied to the solution of two- and threedimensional problems. We recommend the direct method if the Stokes problem has to be solved many times on a given domain. On the other hand, if the Stokes problem is to be solved a few times only or if Nh (the number of boundary nodes) is very large and requires a complicated out of core procedure, we recommend the conjugate gradient method (5.138)-(5.145).
276
VII Least-Squares Solution of Nonlinear Problems
Remark 5.13. The direct method of Sec. 5.7.2.4 for solving the Stokes problem (5.103) appears as a coupling between ordinary finite element methods and those panel methods (discussed in, e.g., Hunt [1]) for solving boundary integral equations.
5.8. Another mixed finite element method for the Stokes and Navier-Stokes problems: applications 5.8.1. Introduction The finite element method introduced in Sec. 5.3 seems unnatural at first glance (despite Remark 5.1), but it has very interesting decomposition properties resulting in efficient direct or iterative Stokes solvers. Actually the above finite element approximation is a variant of a more classical one (advocated by Hood and Taylor [1]) which will be discussed below; we shall also discuss several methods for solving the corresponding approximate Stokes problems. 5.8.2. Formulation of the basic problem: principle of the method Let us consider the following Navier-Stokes problem (with a > 0): au — vAu + (u • V)u + \p = f in Q, V • u = 0 in Q,
(5.153)
u = g on FI with g • n dY = 01. \ Jr ) The incompressibility condition V • u = 0 is equivalent to -uqdx = 0,
VqeH\Q).
(5.154)
From (5.154) it then follows that a possible variational formulation of the Navier-Stokes problem (5.153) is
{u, p) e (H'iQ))2 x L 2 (Q),
u = g on T,
I u • v dx + v Vu • Vv dx + (u • V)u • v dx + \ \p • v dx = Jo Jfi Jn Jn Jci V ve ( H » V • u
2
f • v dx,
,
Vqe H 1 ^ ) .
(5.155)
The mixed finite element approximation of (5.153) which follows is derived directly from (5.155).
5 Numerical Solution of the Navier-Stokes Equations
277
5.8.3. A mixed finite element approximation ofthe Navier-Stokes problem (5.153) We use the notation of Sec. 5.3.3.1; formulation (5.155) clearly suggests the following (well-known) mixed approximation of the Navier-Stokes problem (5.153): F i n d {u,,, ph} e Vgh x Hi such that a Jn
uA • vh dx + v \ \ah • \\h dx + \ (uh • \)uh -\hdx+ Jn Ja
= [ih-yhdx, Jn
\ph • \h dx Jn
Vv,eF O h ,
(5.156)
5.8.4. Convergence results 5.8.4.1. Case of the Stokes problem. Uniqueness of the discrete pressure. The discrete Stokes problem associated with (5.156) is defined by: Find {uh, ph} e Vgh x U\ such that a Jn
uh-\hdx
+v Jn
\ah
\\h dx + \ \ph • \h dx = fh • \h dx, Jn Jn \-ahqhdx
= 0,
VqheHl.
V \h e VOh, (5.157)
From Chapter I, Sec. 7.4.2 it follows that (5.157) has at least one solution but with uh uniquely defined. Actually, if condition (5.119) holds, then {nh, ph} is uniquely defined in Vgh x (Hj,/U). EXERCISE 5.7. Prove that the results of Chapter I, Sec. 7.4.2 can be applied to problem (5.157).
With regard to the convergence of the approximate solutions, it is proved in Bercovier and Pironneau [1] that if the hypotheses on (^) A are those of Lemma 5.1 and Theorems 5.4 and 5.5 of Sec. 5.3.3.3, then the estimates (5.26), (5.27), (5.29), (5.31), (5.32),25 (5.33), and (5.34)25 hold. 5.8.4.2. Case of the Navier-Stokes problem. It has been proved by Le Tallec [3] that the convergence results mentioned in Sec. 5.8.4.1 also hold for the Navier-Stokes problems (5.153), (5.156).
25
With ibh = 0.
278
VII Least-Squares Solution of Nonlinear Problems
5.8.5. Application to the solution of the unsteady Navier-Stokes problem 5.8.5.1. A semidiscrete approximation of the unsteady Navier-Stokes equations. We suppose that the time-dependent problem under consideration is still defined by (5.39)-(5.42); we can then reduce (5.39)-(5.42) to a system of ordinary differential equations which is defined by: Find K ( 0 , ph(t)} e VJt)
x (tf J/R) such that
f - ^ • vh dx + v I Vu, • Vvh dx + I (u, • \)ah -\hdx+
Jn vt
Jn
Jn
f \ph • yh dx
Jn
(5.158)
1
\-ukqhdx
=
uA(0) = aOh e Vh given (uOh: approximation
ofu0).
In (5.158) the space Vgh{t) is defined by (5.45). The approximate problem (5.158) is not directly suitable for computation, and for this purpose a convenient time discretization is required in order to obtain a fully discrete approximate problem; several such time discretization schemes are given in the following sections (actually these schemes are variants of the schemes described in Sec. 5.4 whose notations are retained). 5.8.5.2. An ordinary implicit scheme. This scheme is a variant of scheme (5.46), (5.47) (see Sec. 5.4.1) and is defined as follows ({u^, pi} approximates {a(nk), p{nk)}, where k = At): u ° = u 0 ,, +1
+1
then for n > 0, we obtain u ^ , p?,
f U* +1 ~ n i !
-yhdx
(see (5.158));
(5.159)
from uj| by solving • Vv, dx
f (u"h
dx
Jn Jn vqhenh,
ph
^- T/"+ 1
e v gh
(5.160) where V™h = Vgh(mk), f I = f,(m/c). To obtain {u^ +1 ,pj| +1 } from uj, we have to solve a finite-dimensional nonlinear problem which is a particular case of (5.156) (with a = 1/fe). The above scheme has a time truncation error in O(k) and appears to be unconditionally stable.
5 Numerical Solution of the Navier-Stokes Equations
279
5.8.5.3. A Crank-Nicholson implicit scheme. This scheme (which is a variant of (5.48)-(5.50)) is defined as follows: < = uo*;
(5.161)
then for n > 0, we obtain ujj+ J from uJJ by solving r u| + 1 - u ; _ ^ Jn k
dx
r V n , + 1 / 2 _^
+ v
dx
+
JQ
r (u«+1/2 _ v ) u , , + 1 / 2 . Ja
+ f Vrf +1/2 • vh dx = f f r 1 / 2 • v, dx, Jn
-ul+l'2qhdx
= 0,
Jn
VqheHlh,
{u"h+1/2, p"h+1/2} e V^1'2
Vh dx
Vv, e VOh, x i/i/R (5.162)
^ g Using the relation u^+ 1 = u"h + 2(u£ +1/2 - uj), we can eliminate u^ +1 in (5.162) and therefore reduce this problem to a particular case of (5.156) (with a. = 2/k). The above scheme has a time truncation error in O(k2) and appears to be unconditionally stable. 5.8.5.4. A two-step implicit scheme. This scheme is defined by: < = uoh, uj, given; then for n>\,we
(5.163)
obtain {u^ +1 , p"h+ 1} from u"h, u^" 1 by solving
Uu"h
Jn
v
' u/i
qhax
= yi,
v qhe Hh,
{\\h
, ph
n (5.164) To obtain u/} from u°, we may use one of the two schemes discussed in Sees. 5.8.5.2 and 5.8.5.3. From (5.164) it follows that u^ +1 is the solution of a particular problem (5.156) (with a. = 3/2fe). The above scheme has a time truncation error in O(k2) and appears to be unconditionally stable.
280
VII Least-Squares Solution of Nonlinear Problems
5.8.5.5. Alternating-direction methods. In this section we consider variants of those alternating-direction methods discussed in Sec. 5.4.4. Actually all the comments and remarks made in Sec. 5.4.4.3 also hold for the schemes which follow. A. A Peaceman-Rachford alternating-direction method for solving the unsteady Navier-Stokes equations. This variant of (5.53)-(5.55) (see Sec. 5.4.4.1) is denned as follows (with 0 < 6 < 1): < = u Oh ; then for n > 0, we obtain /• ..11+1/2 _
^—— Jn k/Z
{unh+1/2,
un
pnh+1/2},
a"h
(5.165) +1
from u£ by solving
/.
h
-yhdx
r>
\u"h+1'2-\yhdx+
+ 6v
\p"h+1'2-\hdx
Jn
Jn f f"h+1'2 • yhdx,
+ (1 - 0)v f \u"h• Vv,dx + f (uj• V)u"h-vhdx= (\-u"h+"2qhdx
Jn
= 0,
V v , e Voh,
{K+ \ rf+1/2} £ ^/, + ' / 2 x HlJU,
VqheHl,
(5.166) anrf t/ien r u " + 1 — n"+ 1/2 T -5 —* v, + (1 - 0)v \u"h+l • \yh dx K A Jn I Jn
+ f (u; + ' • V K + ' -vhdx + 0v f V< + 1 / 2 • Vv,rfx+ f Vrf +1/2 • v, ^x Jn
Jn
+1
-v»dx,
Vv.eFc,
Jn
ul + 1 6 ^ + 1 .
(5-167)
B. A second alternating-direction method. This is the variant of (5.166), (5.167) defined as follows (again with 0 < 6 < 1): < = u Oh ; then for n>0,weobtain{unh+m,pnh+1/4},unh /•
Jn
un+
1/4 _
'
,,. K n
un
f V • uj + ^ % Jx = 0, Jn
+ 3l4
,{<+1,pnh+1}fromunhby
f.
" • vfc dx + (1 - 0)v
+ 0v \\n"h-Vvhdx+ Jn
(5.168)
Jn
solving:
/•
V u r 1 / 4 • Vv, dx +
\(u"h-\)u"h-vhdx=\ Jn V q, e H,1,
Jn
Jn
Vrf + 1 ' 4 • v, dx
f"h+ll*-vhdx,
VvteF0l,
{u"/1/4, rf+1/4} 6 V"g;1/4 x Hi/R; (5.169)
5 Numerical Solution of the Navier-Stokes Equations
281
then /.
un
+ 3/4 _
j.n+1/4
—~ k /2
Jn
»
\unh+3l4-\\hdx
\hdx + (h Jn
+ f (uj +3 '* • V K + 3 / 4 • M * + (1 - 0)v f Vu^+1/4 • Vvh dx Jn Jn +3/4
Jn
„
j
V v (= 1/
nn
+ 3/4
p
T/-H+3/4-.
Jn (5.170)
/• i j " + 1
A
i.n + 3/4
^
/•
/•
VuS + 1 -Vv»dx+
vfcdx + ( l - 0 ) v
VpX+1-
+ 0v f V<+3/4 • \\h dx + f (u^ +3/4 • V K + 3 / 4 • vA rfx Jn
Jn
Jn
K + 1 ,K + 1 }eF« h + 1 xHl/U.
(5.171)
5.8.6. Least-squares conjugate gradient solution of the discrete Navier-Stokes problem (5.156) 5.8.6.1. Generalities. The various discrete Navier-Stokes problems that we have encountered in this section are all particular cases of problem (5.156); for this reason we shall study in some detail the least-squares formulation and the conjugate gradient solution of (5.156). Actually the following considerations very closely follow Sees. 5.5 and 5.6. 5.8.6.2. A least-squares formulation of (5.156). A natural least-squares formulation of (5.156) is as follows: Find uh e Vgh such that JHM
< Jh(yh),
V v, e Vgh,
(5.172)
with
Vgh = k £ Vgh, j \ • yhqk dx = 0, A(vft) = ^ f I^K)| 2 dx+U 1
Jn
l
V<j,£ Hi j ,
|V^(v,)|2 dx, Jn
(5.173) (5.174)
282
VII Least-Squares Solution of Nonlinear Problems
where, in (5.174), ^h(\h) ( = %h) is a function of \h through the state equation a K A • Vhdx + v Jn
V ^ • Vi]A dx +
Jn
VTT,, • x\hdx = a Jn
+ v \ \\h • V% dx + I (vh • V)vft • % dx -
f v . ^ f c d x = 0,
v,, • % dx Jn
VqheHl
K , • % dx,
V TJA e F o/ ,,
%hsVOh, nheHl/U.
(5.175)
Jn
We observe that \h is obtained from \h through the solution of a discrete Stokes problem. The above least-squares formulation is justified by the following: Proposition 5.3. Suppose that {uh, ph) e VgH x (Hj,/U) is a solution of (5.156); then uh is also a solution of the least-squares problem (5.175) and, if (5.119) holds, for the corresponding pair {^h, nh} we have \h = 0,
nh = -Ph;
(5.176)
we also have Jh(uh) = 0. 5.8.6.3. Conjugate gradient solution of the least-squares problem (5.172). A. Description of the conjugate gradient algorithm. A conjugate gradient algorithm for solving the least-squares problem (5.172) is defined as follows. Step 0: Initialization u°heVgh given;
(5.177)
then compute z° as the solution of the linear variational equation a f z°h • % dx + v f Vz° • Vt,, dx = < / ; « ) , ifo>, Jn Jn
V t|» e %h, z°h e VOh
(5.178) and set < = zl For m > 0, assuming that u", z™, w™ are known, compute < the following.
(5.179) +1
+1
, z^ , w™+1 by
Step 1: Descent Compute Xm = Argmin J , , « - A < ) ,
(5.180)
XeR
ur i =u^-A m <.
(5.181)
5 Numerical Solution of the Navier-Stokes Equations
283
Step 2: Construction of the new descent direction. Define z™+1 as the solution
of a f z r * •% dx + v f V z r ' • V% rfx = <J;W + *), %>, Jn
Viu 6 Foft;
Jn
(5.182) then (Polak-Ribiere strategy) _ « j n z r 1 • ( # + ' - ff) ^x + v J n \z% + 1 • V(z^ +1 - z?) dx y>n+\
r
i_m|2
JY
i ., f
|\77'"|2jv
'
(5.183) <+1=tf+1
+7m+i<,
(5.184)
m = m + 1 go to (5.180). B. Calculation ofJ'h(u% + 1) and z" + 1 : further comments. A most important step in order to use algorithm (5.177)-(5.184) is the calculation of J^(u™+1); another important step is the calculation of z™+1. Due to the importance of these steps, we shall give the corresponding calculations in detail. We have 5J = <J'h(Yh\ 5yh) ==aa ff^^ ••S^dx S^dx J Jn
++vvf f V^ • \S^ dx; JJn
(5.185)
to express b\h as a function of 5\h, we observe that (5.175) has the following variational formulation: l,h e VOh and V % e VOh, we have
C
C
C
a \ %h-t\hdx + v \ \Z,h-\r\hdx Jn Jn
C
= a \ vh-r\hdx + v \ \vh • V% dx Jn Jn
f-iudx. EXERCISE
(5.186)
5.8. Prove the equivalence between (5.175) and (5.186).
Relation (5.186), in turn, implies 5Z,h e Voh and V % £ VOh, we have a \ d^h • r\h dx + v \S$h •\r\hdx Jn Jn + Uovh-\K-r\hdx+ Jn
= ct
Jn
\(yh-Y)5vh-r\hdx. Jn
d\h • x\k dx + v
Jn
\S\h • \r\h dx (5.187)
284
VII Least-Squares Solution of Nonlinear Problems
Since \h e VOh, we can take % = \h in (5.187); from (5.185) it then follows that SJH = <^(Vfc), <5vh> = a f %h • 5vh dx + v f V$» • V<5v, dx Jn Jn +
f (5v» • V)vfc • ^ dx + f (vA • V)<5v, •!;,, dx. Jn
(5.188)
JCl
Finally from (5.188) it follows that J'h(\h) can be identified with the linear functional on VOh defined by dx + ff (r\hh • \)\ \)\h • ^ dx
f (v, • V ) % • \h dx.
+
(5.189)
From (5.189) it then follows that to compute z™+1 from u™+1, we first have to solve the discrete Stokes problem (5.175), (5.186) (with \h = u™+1) to obtain ^ + 1 ; then from (5.182), (5.189) we obtain z^ +J via the solution of
a f z^ + 1 • % dx + v f V z r ! • V% dx = a f ^ Jn
Jn
• n, dx
Jn
+ v f V ^ + 1 • V % dx + f ( % • V K + 1 • ^ Jn
+ 1
+ 1
dx
Jn
+ fK+1-V)%-^+1dx. (5.190) Jn Actually (cf. Exercise 5.8) problem (5.190) is equivalent to the discrete Stokes problem a ( z%+l-r\hdx + v fvz^-Viifcdx + f V < + 1 -%dx Jn Jn Jn = a f ^ + 1 - i h d x + v fv§J' +1 -Vi h dx+ Jn
Jn
+ ("(ur'-V^-^^dx, Jn
fv-z^+1^dx = 0, Jn
\(n,-VM+1-tS:+1dx Jn
VikeFOi, V^efli,
{zr>: + 1 }e^xHl/K; (5.191)
+1
co™ is clearly a (discrete) pressure. To summarize, at each iteration of the conjugate gradient algorithm (5.177)-(5.184), we have to solve several discrete Stokes problems, namely:
5 Numerical Solution of the Navier-Stokes Equations
285
(i) the discrete Stokes problem (5.175), (5.186) with vh = < + 1 to obtain ^+1from<+1; (ii) the discrete Stokes problem (5.182), (5.190), (5.191) to obtain z£"+1 (iii) the solution of the one-dimensional problem (5.180) requires several calculation of Jh, i.e., several solutions of the state equation (5.175), (5.186). Thus, efficient solvers for the discrete Stokes problem (5.157) will be important tools for solving the Navier-Stokes equations by algorithm (5.177)(5.184) via the least-squares formulation (5.172); actually these solvers will also be useful for solving the Stokes parts of the alternating-direction algorithms (5.165)-(5.167) and (5.168)-(5.171) of Sec. 5.8.5.5. 5.8.7. On numerical solvers for the discrete Stokes problem (5.157) 5.8.7.1. Synopsis. From Sees. 5.8.5 and 5.8.6 it follows that efficient solvers for the discrete Stokes problem (5.157X i.e.: Find {uh, ph} £ Vgh x Hi such that a \ uh • \h dx + v \ \nh • \\h dx + \ \ph • \h dx Jn Jn Jn = \th-\h dx, V vh e VOh, V • uhqh dx = 0, V qh e Hj,, Jn Jn will be advantageous if one wishes to solve the steady and unsteady NavierStokes equations through the finite element approximation of Sec. 5.8.3. In the following sections we shall discuss the solution of (5.157) by either direct or iterative methods. 5.8.7.2. Direct solution of (5.157). We are going to give a linear system equivalent to (5.157); we recall that Hi = {qh I qh e C°(Q), qkiTePu
V T e STh}.
We now define Hi by either Hi = {zh\zh e C°(Q), zh]Te P2,
V T e ^ } if Vh is defined by (5.18), (5.192)!
Hi = {zh | ih G C°(Q), z h|T e Pu
V T e ^ } if Vh is defined by (5.21).
or
(5.192)2 We clearly have Vh = Hi x Hi; we also have Voh = H\h x Hlh, where H^ = Gln
H&a) = {zh\zh e Hi zh\r = 0}-
(5-193)
286
VII Least-Squares Solution of Nonlinear Problems
We now define Zh = {P e Q, P vertex of T e &~h}; we have d\m(Hl) = Card(EA), and we define Mh by Mh = dim(Hl). Suppose that ~Lh = {Pj}fi"i; we associate to Sft the vector basis of H\ defined by . Fori =
(5.194)
l,...,Mh,
w, e Hi, wiP^ = 1, HPj)
= 0,
V j / i,
l<j<
Mh
(5.195)
We then have Mh
«*=E«*(^M,
V^eH, 1 .
(5.196)
;= i
We similarly define I,, = {g e Q, g vertex of T e &~k, or g midpoint of an edge of T e ,fh} = {Q e U, Q vertex of T e .fh}, (5.197) I Oh = { P e I A , P ^ r } ,
(5.198)
JV, = Card(£ft) = dim(^),
(5.199)
NOh = Card(£ 0A ) = dimiHik).
(5-200)
We suppose that £
, ^
+ 1
(5.201) ,
(5.202)
and we associate to f.oh, t.h the following vector bases of Hlh and Hi, respectively: @OH = {«;}^i,
(5-203)
£* = {«,-}£!
(5-204)
with w;-defined by: V j = 1 , . . . , Nh, we have w, e H,1, wJiQj) = 1, w / g , ) = 0,
Vfe^ j . (5.205)
We then have V \h e Vh (resp., Vv ft e Voh), vft = t vh(Qj)wj
(5.206)
V* = ^v»(6j)*i).
(5-207)
(resp.,
5 Numerical Solution of the Navier-Stokes Equations
287
Let {ah, ph} be the solution of (5.157). We have u, = funiQiWj
+
t
SH(QJ)WJ;
(5-208)
u/, = {UIH, u2h) (with ulh, u2h e Hi),
(5.209)
g* = {
(5-210)
if
it follows from (5.208) that we have Vr=l,2
urh= furh(Qj)wj+ j=i
t
grh(Qi)Wj.
(5.211)
j=JV 0h +l
Similarly we have Mh
PH= E P ^ K -
(5.212)
Using the above relations, the discrete Stokes problem (5.157) is equivalent to the following linear system of 2NOh + Mh equations: £ (a f wjwt rfx + v f \wj • Vw; dx)ulh(Qj) j=i\
Jn
Jn
= f AhWi dxJn
I
f
(a f WjWt dx + v \\wj- \wt
j=NOh+i\
Vi,
Jo
Jo
dx)glh(Qj), I
l
(5.213)!
f (a f WjWi dx + v f Vw; • Vw;
k=
c
^ / r
r
\
V i, l < i < NOh,
(5.213)2
V /, 1 < / < Mh,
(5.214)
288
VII Least-Squares Solution of Nonlinear Problems
whose unknowns are {{ulh(Qj), «2»(Qj)}}J'™i, {ph{Pk)}^
(and where fh =
Let us define UA e R2Noh and Ph e UMh by U* = {{ulh(Qj\ u2h(Qj)}}pu
P , = {ph(Pk)}kLhil
(5-215)
since Wj e Hl(Q), Vy = 1 , . . . , iV0A, we have (Green's formula) 5w-
f „ dwl
r-^ wtdx = —
v'-l
Wj- —— ax,
N ° vT-1
M
(5 216)
'
Multiplying the equation (5.214) by - 1 , it follows from (5.215), (5.216) that the linear system (5.213), (5.214) has the following block representation: AhVh + B^P, = V B,U, = ch,
(5.217)
where Ah is a 2NOh x 2NOh symmetric positive-definite sparse matrix and where we have (from (5.214)) ch e R(Bh); such linear systems have been considered in Chapter I, Sec. 7.4.2. The matrix associated with (5.217), i.e.,
is symmetric, indefinite, and sparse; we actually have Ker(ja/ft) = {0} x Ker(B^), and if (5.119) holds, we have Ker(BD = {Qh | Q, = {c}^\,
c e R},
(5.218)
i.e., dim Ker(BJ,) = 1 o R a n k ^ , , ) = 2NOh + Mh - 1. From (5.218) it then follows that we can trivially reduce (5.217) to a system whose matrix is nonsingular (and still symmetric, sparse, and indefinite) by taking (for example) the MAth component of Ph equal to zero. The direct solution of systems like (5.217) (by sophisticated variants of the Gauss elimination method) has inspired several authors; let us mention among them Parlett and Reid [1]. Bunch and Parlett [1], Aasen [1], Bunch and Kaufman [1], Barwell and George [1], Duff, Munksgaard, Nielsen, and Reid [1], and Duff [1]; see also the references therein. The methods originating from the above references are quite storage demanding, which can be a major drawback when trying to solve very complicated flow problems (actually these methods can also be used—theoretically at least—for solving the discrete Stokes problem (5.113) of Sec. 5.7.2.1). The solution of (5.157) by iterative methods (less storage demanding than the above direct methods) will be discussed in the following sections.
5 Numerical Solution of the Navier-Stokes Equations
289
5.8.7.3. Iterative methods for solving the discrete Stokes problem (5.157) 5.8.7.3.1. Generalities: synopsis. In Sec. 5.8.7.2 we have shown that the discrete Stokes problem (5.157) is in fact equivalent to a linear system (namely (5.217)) having the following very special structure (we have slightly modified the notation of Sec. 5.8.7.2 and also of Chapter I, Sec. 7.4.2): Ax + B'X = b, Bx = c,
(5.219)
where A is an JV x N symmetric positive-definite matrix, B is an M x N matrix, b e UN, c e R(B). From Chapter I, Sec. 7.4.2 it follows that (5.219) has a unique solution {x, X} in UN x R(B); it also follows from Chapter I, Sec. 7.4.2 that {x\ X'} eUN x UM is a solution of (5.219) if and only if x' = x,
X' = X + \i, n e Ker(B').
(5.220)
The linear system (5.219) can also be written (A
BVx
B 0/w
W
(5 221)
-
since the matrix in (5.221) is symmetric and indefinite, this linear system can be solved iteratively by those Lanczos methods (generalizing the standard conjugate gradient methods) introduced in Lanczos [1] and some improved versions of which are discussed in Paige and Saunders [1] and Parlett [1]. Actually, from our numerical experiments26 these Lanczos methods seem quite delicate to use and in the following sections we shall concentrate on iterative methods taking into account the fact that A is a symmetric and positive-definite matrix. More precisely, in Sec. 5.8.7.3.2 (resp., 5.8.7.3.3) we shall consider the solution of (5.219) by an Uzawa conjugate gradient algorithm (resp., by an augmented Lagrangian method); then in Sees. 5.8.7.3.4 and 5.8.7.3.5 we shall apply these methods to the solution of the discrete Stokes problem (5.157). Finally, in Sec. 5.8.7.4 we shall discuss the solution of the continuous Stokes problem au - vAu + Vp = finQ,
V • u = 0 in Q, u = g on F
(5.222)
by the continuous analogues of the algorithms of Sees. 5.8.7.3.2 and 5.8.7.3.3. 5.8.7.3.2. Iterative solution of (5.219) by an Uzawa conjugate gradient algorithm. From (5.219) it follows that X is a solution of the linear system, in RM, BA" WX = BA" xb - c. ' On discrete Stokes problems precisely.
(5.223)
290
VII Least-Squares Solution of Nonlinear Problems
We clearly have Ker(BA- 1 B') = Ker(Bf), 1
(5.224)
t
which, since the matrix BA~ B is symmetric, implies that R(BA" 1 B') = R(B).
(5.225)
-1
Since c and B A b belong to R(B), it follows from (5.225) that (5.223) has a solution; in fact it has a unique solution X in R(B) since B A ^ B ' is an isomorphism from R(B) onto R(B). Let us define 33 and P by P = BA~xb-c;
0£ = BA ^ B ' , we then have
03X = p.
(5.226)
Since 3B is symmetric and positive semidefinite and since P e R(B), we can solve (5.226) (i.e., (5.219)) by the following conjugate gradient algorithm (where (•, •) denotes the usual scalar product of UM and | • | denotes the corresponding norm). Step 0: Initialization X° e RM arbitrarily given;
(5.227)
g° = 0SX° - p ,
(5.228)
w° = g°.
(5.229)
Then for n > 0, we obtain I " + l , g * + 1 , « " + 1 from X", g", w" by the following. Step 1: Descent
p =
" ^^)=lShrr r
+1
= r -pnw.
Step 2: New descent direction g" +1 = mX"+1 - p ,
(5 230)
-
(5.231) (5.232)
lgn+1,2
y =
" \¥\2 '
w"
+1
= g"
+1
+ ynw",
(5 233)
'
(5.234)
n = n + 1 go to (5.230). From (5.231), (5.232) it follows that we also have g"
+1
= g" - p n ^w",
which we should use instead of (5.232) in practice. Concerning the convergence of algorithm (5.227)-(5.234) we have:
(5.235)
5 Numerical Solution of the Navier-Stokes Equations
291
Proposition 5.4. Let Mo be the dimension of R(&)21 ( = R(B)); then there exists n0 < Mo such that X"° is a solution of'(5.223), (5.226); more precisely we have X»° = % + 1°,
(5.236)
where 1% is the component ofk° belonging to Ker(B') in the decomposition UM = R(B) © Ker(B'). EXERCISE
5.9. Prove Proposition 5.4.
Hint: First prove that (g',gj)
= 0,
Vi,;,
/#;,
(g ; ,w0 = 0,
Vi,;,
i>j,
Vi,;,
i#;,
;
J
0&w , w ) = 0, and use relation (5.235).
Remark 5.14. If 1° e R(B) (k° = 0 for example), it follows from Proposition 5.4 that algorithm (5.227)-(5.234) converges to i in a finite number ( < M 0 ) of iterations. Remark 5.15. The above finite termination result holds if roundoff errors are not considered. However, even with roundoff errors, a practical convergence can be obtained in much less than M o iterations, due to the superlinear convergence of conjugate gradient algorithms (see Polak [1] and Hestenes [2] for more details on the convergence of conjugate gradient methods). For M o sufficient large, numerical experiments show a practical convergence in 0{^/M0) for most applications (this is the case, in particular, for algorithm (5.138) (5.145) in Sec. 5.7.2.5.). Practical implementation of algorithm (5.227)-(5.234). In practice we should use (5.227)^(5.234) as follows. Step 0: Initialization 1° e R M ,
(5.237) 0
x° = A - ^ b - B ^ ) , g° = c - Bx°, Then forn>0
we define Xn+\gn+1,Y/n+l
Mo = rank of B
w° = g°. from V, g", w" as follows.
(5.238) (5.239)
292
VII Least-Squares Solution of Nonlinear Problems
Step 1: Descent ^ = A-1BV,
(5.240)
if = B£" (= ^w"),
(5.241)
2
ft,=
Is" I
'. '
Xn+1 = V -
,
(5.242) n
Pnyv
.
(5.243)
g" +1 = g" - A,tf,
(5.244)
Step 2: New descent direction
'•-ww «+i
_ g»+i + ynW";
(5245) (5.246)
n = n + 1 go to (5.240). Once {^"}«>0 has converged to the X solution of (5.226) (and (5.223)), we obtain the corresponding x in (5.219) via the relation x = A - 1 ( b - B'X).
(5.247)
From (5.237)-(5.246) it follows that each iteration of algorithm (5.227)(5.234) will require the solution of a linear system whose matrix is A; since A is symmetric and positive definite, one may use a Cholesky factorization A = LL' of A (with L lower triangular) done once and for all; each iteration of (5.237)-(5.246) will therefore require the solution of two triangular linear systems, which is not a very costly operation. Relation with Uzawa algorithms. We would now like to justify the title of this section (Sec. 5.8.7.3.2); to achieve this goal, let us consider the solution of (5.223), (5.226) by a standard gradient method with a constant step p (> 0); such an algorithm is written as follows: X°eUM;
(5.248)
then for n > 0 we define X"+1 from X" by X»+i
=
xn - p(mXn - P).
(5.249)
Since 3d = BA~ 'B', (5.249) is in fact equivalent to x" = A - 1 ( b - B ' l . " ) ,
(5.250)
r + 1 = JL" + p(Bx" - c).
(5.251)
In (5.248), (5.250), (5.251) we recognize an Uzawa algorithm (see, e.g., G.L.T. [3, Chapter 2]) applied to the solution of the linear system (5.219); algorithm (5.227)-(5.234) (in its equivalent formulation (5.237)-(5.246))
5 Numerical Solution of the Navier-Stokes Equations
293
is clearly a sophisticated variant of the Uzawa algorithm (5.248), (5.250), (5.251). Algorithm (5.248), (5.250), (5.251) is a particular case of a general family of algorithms for computing the saddle points of Lagrangian functional; in the present problem the Lagrangian functional is defined by J2?(y, n) = i((Ay, y)) - ((b, y)) + (n, By - c)
(5.252)
(with ((•, •)) the usual scalar product of UN), and the associated saddle-point problem is as follows: Find {x, )i}elR"x UM such that if(x, n) < <£(x, X) < i?(y, X),
V {y, |i} E UN X R M .
(5.253)
As usual (see G.L.T. [1], [3], Cea [1], [2], Rockafellar [1], and Ekeland and Temam [1] for more details) the primal problem associated with (5.253) is defined by Inf Sup J??(y, p)
(5.254)
Sup M £C(y, p).
(5.255)
and the dual problem by
We then have: Proposition 5.5. The primal problem reduces to
xeH,
J(x) < J(y),
VyeH,
(5.256)
where H = {y|y E UN, By - c = 0}, J(y) = K(Ay, y)) - ((b, y)),
(5.257) N
Vy e U ;
(5.258)
the dual problem is defined by X e UM,
J*(X) < J*Qi),
V fi e RM,
(5.259)
where j*(u) = i(BA~ iB'fi, fi) - (BA" J b - c, ji)
(5.260)
(i.e., the dual problem is the linear system (5.223), (5.226)). PROOF. From (5.252), (5.258) we have Sup i?(y, |i) = J(y) + Sup (n, By - c);
(5.261)
we clearly have (5.262)
294
VII Least-Squares Solution of Nonlinear Problems
From (5.262) it follows that the functional y -> Sup (n, By - c) is in fact the indicator functional IH of the affine subspace H; since c e R(R) implies that H # 0, we have lH convex, proper, and l.s.c.. Hence the primal problem reduces to Inf {J(y) + IH(y)},
(5.263)
yel"
which is equivalent to (5.256) (from Chapter I, Sec. 2). Now consider the minimization problem Infi?(y,n);
(5.264)
ye«N
it has a unique solution x(n) which is also the solution of the linear system Axfti) = b - B'n.
(5.265)
Inf £C(y, (t) = -KBA" >B'n, |i) + (|i, BA" 'b - c) - K(b, A" 'b)),
(5.266)
From (5.252), (5.265) we then have
which implies that the dual problem (5.255) reduces, in fact, to (5.259).
•
5.8.7.3.3. Iterative solution of (5.219) by an augmented Lagrangian algorithm.
This section closely follows Fortin and Glowinski [3]. Principle and description of the method. We complete Remark 7.5 of Chapter I, Sec. 7.4.2. Let r be a positive parameter; we observe that the linear system (5.219) is in fact equivalent to Ax + rB'(Bx - c) + B'l = b
Bx = c,
(5.267)
and that (5.267) characterizes {x, k} as a saddle point over UN x IRM of the (augmented) Lagrangian functional £Cr: UN x IRM -> IR defined by if£y, p) = if(y, ji) + Y- | By - c | 2 ,
(5.268)
with if(-, •) defined as in (5.252). Applying the standard Uzawa algorithm to the saddle-point problem: Find {x, 1} e UN x UM such that &£*,
|i) < JSf/x, X) < jSf r (y, 3L),
V {y, n } € R
N
x
RM,
(5.269)
we obtain:
X°eUM given;
then for n>0we define x", V+1 from I" by:
(5.270)
5 Numerical Solution of the Navier-Stokes Equations
295
Find x" e UN such that JS?,(x", X") < i f / y , X"),
VyeR*,
X" +! = r + p(Bx" - c);
(5.271) (5.272)
actually (5.271) is equivalent to the linear system (A + rBB)x" = b + rBc - B'X".
(5.271)'
Convergence of algorithm (5.270)-(5.272). Concerning the convergence of algorithm (5.270)-(5.272), we have: Proposition 5.6. We have convergence of algorithm (5.270)-(5.272), V X° e (RM, if and only if
0
(5.273)
w/zere AMo is the largest eigenvalue of A" ^ ' B . / / (5.273) holds, then lim {x", X"} = {x, X + X°2},
(5.274)
n—y + GO
w/iere {x, X) is the unique solution of (5.219) in UN x R(B) and X° is the same as in the statement of Proposition 5.4. PROOF. Let us define x", X" by x" = x" - x,
I" = V - (i + l°2).
(5.275)
M
We have R = R(B) 0 Ker(B'), and
r = x."1 +r 2 , 6
where X" e R(B). A" Ker(B') (the above decomposition is unique). From (5.272) it follows that W+l -We R(B), which shows that X\ = \%, Vn,
1"6R(B),
V«.
(5.276)
By subtraction we obtain, from (5.267), (5.271)', (5.272), (A + rB'B)x" + B l " = 0, X"
+1
=X" + pBx".
(5.277) (5.278)
Combining (5.277) and (5.278), by elimination of x" we obtain A-'BT"-" = X~lWl" - pA-'B'B^ + rA-'B'By'A-'Wl".
(5.279)
Let us define {z"}n>0 by z" = A"-JB'X"; it follows from (5.276) that z"6/?(A- 1 B'B),
Vn,
(5.280)
and from (5.279) that z" + ! = (I - pA"'B'Bil + rk~'B'B)" ^z".
(5.281)
296
VII Least-Squares Solution of Nonlinear Problems
Figure 5.3 Since i?(B) and R(A~ 'B'B) are isomorphic (via A" 'B'), the convergence of z" to 0 will imply (5.274) (with the same speed of convergence). The eigenvalues of A~'B'B are non-negative; let {A;}?!0! be the strictly positive eigenvalues of A~ 'B'B, ordered such that 0
v
Mo-
From the properties of A and B'B, there exists, a vector basis {Wj}1L\ of R(A ^ ' B ) such that w, is an eigenvector of A^B'B associated with A,, V ; = 1,..., M o , ((Aw;, w7)) = 0,
V 1,7,
Since z" 6 R(A~ ^'B), we have z" = YPi
1 < i,j <M0,
iyt j .
(5.282)
z
>h and from (5.281)
V n and V i = 1,..., Mo, we have (5.283)
1 + rA,-
Relation (5.283) implies that we have convergence of z" to 0, V >,° e RM, if and only if 1 + (r - p)A,
1+
V i = 1,
...,M0.
(5.284)
We have (5.284) if and only if p obeys (5.273); this clearly appears from Fig. 5.3, where we have shown the graphs of the functions - p)A, for several A,.
As a corollary of Proposition 5.6 we have:
•
5 Numerical Solution of the Navier-Stokes Equations
297
Proposition 5.7. Suppose that p = r; then for r sufficiently large the convergence ratio of algorithm (5.270)-(5.272) is of order 1/r. PROOF. The above result follows directly from (5.283) which shows that if P = r, then V n and V i = 1 , . . . , M o , we have 1 + rA
Remark 5.16. From (5.283) (and Fig. 5.3) the optimal value of p is given by Popt = r +
KAi + AMo) + 2 2rA1AMo + (Aj + A M o ) '
in fact if r is sufficiently large, our numerical experiments show that there is no practical gain in using p = popl instead of p = r. Remark 5.17. In Fortin and Glowinski [3] one will find a conjugate gradient variant of algorithm (5.270)-(5.272) (reducing to algorithm (5.237)-(5.246) if r = 0); actually if p = r and r is sufficiently large, algorithm (5.270)-(5.272) has a very fast convergence and does not require a conjugate gradient acceleration. Remark 5.18. Theoretically, the optimal strategy on r is to take this parameter as large as possible; in practice there is an upper limit due to roundoff errors (see Chapter VI, Sec. 4.3 for further comments). Actually some authors advocate the use of algorithm (5.270)-(5.272) with a sequence of parameters {rn)n>o > obeying lim n ^ + oorn = + co, instead of a constant parameter r; in the particular case of algorithm (5.270)-(5.272), this strategy could be extremely costly since each iteration would require the solution of a linear system whose matrix depends upon n. 5.8.7.3.4. Application of the conjugate gradient algorithm (5.227)-(5.234) to the solution of the discrete Stokes problem (5.157). Instead of (5.227)-(5.234) we shall use its equivalent formulation (5.237)-(5.246). The courageous (or skeptical) reader will verify that we obtain an algorithm whose variational description is given by the following. Step 0: Initialization p° 6 Hi arbitrarily given;
(5.285)
compute u° e Vgh such that
a \u°h-yhdx •in
+ v f v n ? . V v » d x = [th-yhdxJQ
Jn
fvpj?-v f c dx,
Jn
Vv,eF 0 / ,; (5.286)
298
VII Least-Squares Solution of Nonlinear Problems
compute z° e Hj, such that (z°z°hqhqhdx = \\-n°hqhdx, Jo Jn w°h=z°h. Then for n > 0 compute p"h+1,
z"h+1,wnh+i
V^eH,1,
(5.287) (5.288)
from p"h, z"h, wnh as follows
Step 1: Descent Compute xl « f tk • vhdx Jn
e
Voh such that
+ v f \fh Jn
• Vvft dx = -
f VwJ • v, dx, Jn
V v, e F o , ;
(5.289)
then _
Jn PS+1=P!!-P»W|!.
(5.291)
2: Afew descent direction Compute zl+1 e H^ such that
f z j + ^ , rfx = f zXfc dx - p B [ V • zX9h dx, Jn
Jn
V qh e H,1,
(5.292)
Jn
(5-293) <
+1
=znh+> +ynw"h
(5.294)
n = n + 1 go to (5.289). Thus each iteration of (5.285)-(5.294) costs: (i) two uncoupled discrete Dirichlet problems 28 (associated with al — vA) to obtain fh from w"h via (5.289); (ii) the linear problem (5.292) to obtain z\+ x from z\ and fh. Remark 5.19. Algorithm (5.285)-(5.294) is a discrete variant of algorithm (13.17)—(13.27) in Glowinski and Pironneau [3] (see also Sec. 5.8.7.4, where the solution of the continuous Stokes problem (5.222) by a continuous analogue of (5.285)-(5.294) will be discussed).
28
Three in dimension 3.
5 Numerical Solution of the Navier-Stokes Equations
299
Remark 5.20. We can avoid the solution, at each iteration, of the linear problem (5.292) by using, over Hi, instead of the L2(Q)-scalar product, the following scalar product 29 Mh
(yh, qh\ = E yh(Pdqh{Pd,
V yh, qh e Hi,
(5.295)
i= 1
where (cf. Sec. 5.8.7.2) {Pjji*! is the set of the vertices of Th. If one uses the scalar product (5.295), one would have to replace (5.287), (5.290), (5.292), (5.293) by z°heHl,
(z°h,qh)h = f V • u°hqh dx,
VqkeH},,
(5.287)'
(~n .,,n\ w V-hi
z"h+1e
Hi
h)h
(.-r \ qh\ = (rj, qh\ -
Pn
f V • jfi,qh dx,
V
qheHi, (5.292)'
f-n+l
U
-n z
respectively. To be more precise, from (5.292)', (5.295) we have z"h+ \Pt)
= zliPt)
- p
H
\ \ - x>i dx,
V i = 1 , . . . , Mh,
(5.296)
where w; is the j'th element of the vector basis 3%h of Hi denned by (5.194), (5.195). 5.8.7.3.5. Application of the augmented Lagrangian algorithm (5.270)-(5.272) to the solution of the discrete Stokes problem (5.157). There are in fact several possibilities for implementing algorithm (5.270)-(5.272), depending upon the scalar product equipping Hi; a most natural choice is the following variant of (5.295): (}'„, qh)h = I coiyh(P,)qllLPd,
V yh, qh e Hi,
(5.297)
where {Pi}f=i is the set of the vertices of &~h and ii =
Jn (with W; as in (5.296)); we have, in fact,
w; dx
a;,- = | meas ' We can also use scalar products other than (5.295) (like (5.297) in Sec. 5.8.7.3.5, for example).
300
VII Least-Squares Solution of Nonlinear Problems
where O ; is the union of those triangles of 2Th having P, as a common vertex (we may consider that (yh, qh)h is obtained from j n yhqh dx through an approximate numerical integration procedure, namely the two-dimensional trapezoidal rule). Now, if \h e Vh, we define the approximate divergence operator div,, :Vh-+Hj, by Mh
l / r
div,v,= X -
\ V-vhWjixk;
(5-298)
the approximate incompressibility condition
L
is then equivalent to
Jn \-vhqhdx
= 0,
VqheH\,
div, v, = 0.
(5.299)
The augmented Lagrangian algorithm (5.270)-(5.272) applied to the solution of the discrete Stokes problem (5.157) then takes the following form (with r > 0, p > 0): p° e Hi arbitrarily given; then for n > 0, we obtain uJJ and p"h
+l
(5.300)
from pjj by the following:
Compute ujj £ Vgh such that a \u"h-\hdx Jn
+ v | V< • \\h dx + r(div» u"h, divh \h)h Jn
= [ih-ihdxJn
{\pnh-vhdx, Jn
Vvh6FOh;
(5.301)
then Pnh+1=pnh-
Pdivhu"h.
(5.302)
A good choice for p is (still) p = r. Thus each iteration of algorithm (5.3OO)-(5.3O2) costs the solution of the linear problem (5.301) which is clearly related to the continuous operator 30 v -» a\ - vAv - rV(V • v),
(5.303)
which is a variant of the linear elasticity operator. We have to observe that (unlike the operator v -> av — vAv) the operator (5.303) couples the components of v (through the term — rV(V • v)); a similar property holds for the 30
Actually this operator also plays an important role in the compressible Navier-Stokes equations.
5 Numerical Solution of the Navier-Stokes Equations
301
discrete variant of the operator (5.303) occurring in (5.301). Such coupling makes the solution of (5.301) much more costly than, for example, the solution of (5.289) in algorithm (5.285)-(5.294) of Sec. 5.8.7.3.4. Remarks 5.16, 5.17, and 5.18 of Sec. 5.8.7.3.3 still hold for algorithm (5.300)(5.302). A continuous analogue of algorithm (5.300)-(5.302) will be discussed in Sec. 5.8.7.4. Some variants of the above methods for solving the Stokes and NavierStokes problems can be found in Fortin and Thomasset [2], Thomasset [1], and Segal [1]. 5.8.7.4. Some remarks on the iterative solution of the continuous Stokes problem (5.222) 5.8.7.4.1. Motivation. Suppose that an infinite-dimensional problem (P) has been approximated by a finite-dimensional problem (P h ); also suppose that a given algorithm for solving (Ph) is in fact the discrete analogue of an algorithm for solving (P). In many cases the convergence of the infinite-dimensional algorithm will imply the convergence of the discrete one in a number of iterations independent (or "almost" independent) of the discretization parameter h; it is this observation which has motivated our study of the continuous analogues of the algorithms of Sees. 5.8.7.3.2 and 5.8.7.3.3. The following analysis closely follows Glowinski and Pironneau [3, Sees. 13.3.2, 13.3.3]. 5.8.7.4.2. A fundamental theoretical result. From now on, Q is a bounded domain of UN (N = 2, 3 in practice) and F ( = <3Q) is regular (say Lipschitz continuous). We define H c L2(Q) by r
q(x) dx = ! The convergence of the iterative methods, to be described in the sequel, will follow from: Theorem 5.10. Let d: L2(Q) ^ L2(Q) be defined by q e L2(O), av — vAv = —\qin Q, ve(/fJ(Q)) JV
(which implies v| r = 0),
= V • v.
(5.304) (5.305) (5.306)
Then stf is H-elliptic, self-adjoint, and automorphic from H onto H (i.e., 3 fi > 0 such that (.siq, q)L2 > P\\q\\2L2, V q e H). The proof of Theorem 5.10 can be found in Crouzeix [2]. Remark 5.21. The discrete forms of srf are, in general, full matrices.
302
VII Least-Squares Solution of Nonlinear Problems
5.8.7.4.3. Gradient methods and variants. Let us define H ' c: (H^Q))" by 31 H* = {v | v e (H1(^))JV, v = g on T}.
(5.307)
2
Now let {u, p} e H* x L (Q)/IR be the solution of the Stokes problem au - vAu + Vp = finQ,
V • u = 0 in Q, u = g on T,
(5.308)
and u0 be the solution of the (vector) Dirichlet problem au0 — vA0 = f in Q, u0 = g on P.
(5.309)
Problem (5.309) has a unique solution, and by subtracting (5.309) and (5.308), we have oc(u - u0) - vA(u - no) = - Vp in Q, u - u0 e ( H 0 W Hence (from Theorem 5.10), sdp = V • (u — u0) = — V • u 0 . In other words, the pressure is the unique solution, in L2(O)/IR, of sip = - V - u 0 .
(5.310)
Due to the properties of si (see Theorem 5.10), it is natural to solve (5.310) (and therefore the continuous Stokes problem (5.308)) by iterative methods such as the method of steepest descent (or even better, by a conjugate gradient method). Gradient method with fixed step size. For a given p > 0, consider the following algorithm: p° e L 2 (Q) given arbitrarily;
(5.311)
then for n > 0, p" given, compute p « + 1 = p " - p ( ^ p " + V-u 0 ).
(5.312)
In practice one has to replace (5.312) by au" - vAu" = f - Vp" in Q, u" = g on T, p»+i
= p
" _ pv-u".
(5.312a) (5.312b)
Remark 5.22. To solve (5.312a), we have to solve N independent Dirichlet problem for al — vA (in practice, N = 2, 3). Remark 5.23. The previous method is close to the artificial compressibility method of Chorin and Yanenko. We now discuss the convergence of algorithm (5.311), (5.312); we have, indeed, the following result:
We suppose that g e (H 1/2 (n))' v with J n g • n dT = 0.
5 Numerical Solution of the Navier-Stokes Equations
303
Theorem 5.11. If in (5.311), (5.312) we have 0 < p < 2 ^ ,
(5.313)
then, V p° e L 2 (fi), we have in (H^Q))* x L2(Q), strongly,
lim {u", p"} = {u, p)
(5.314)
vvftere {u, p} is the solution of the Stokes problem (5.308) with \np dx = } n p° dx. Moreover, the convergence is linear.32 PROOF. It suffices (from (5.312a)) to prove the above convergence results for the sequence (p"}n>o> from (5.312b) we have f C C C C \ p" dx "Si
p " +1 dx = p
V • u" dx = p
** a
u" • n dY = *^r
^JJ
g • n dT = 0 *r
which implies \ p" dx = \ p° d x ,
V n > 1.
(5.315)
Now consider that solution {u, p} of the Stokes problem (5.308) such that \pdx= and define u" (e (H^Qf)
\p°dx,
(5.316)
and p" by
u" = u" - u, then from (5.315), (5.316) we have p"eH,
pn = p" - p; V n.
(5.317)
We also have (from (5.310)) p = p — p(jrfp + V • u0).
(5.318)
Subtracting (5.318) from (5.312) (resp, (5.308) from (5.312a)), we obtain pn+l _ pn _ pjtfp"
(5.319)
(resp, u" e (Hl0(Q.))N,
au" - vAu" = - Vp" in Q).
(5.320)
Let us denote || • || L 2 (n) by |-|; from (5.319) we have
We actually have (from (5.320)) ,a/p" = V • u"
(5.322)
from which it follows that p", p") L 2 (n) =
:
V • u"p" dx = -
i.e., at least as fast as the convergence of a geometric sequence.
(5.323)
304
VII Least-Squares Solution of Nonlinear Problems
(where < •, • > denotes the duality between (H£(£!))" and (H~ \Q.))N); it then follows from (5.320), (5.323) that (j*?, p")L2(n, = « I" |B"|2 dx + v f | W f dx •>ci
•'n
which, combined with (5.321), implies that \p"+l\2 = |p"| 2 - 2p(a f |u"|2 dx + v f |V0"|2 a!x) + p 2 f |V • u"|2 dx. (5.
324)
Let v e (i/ 1 (n)f; we have (if v = {y;}?= t ) |V-v|2rf.Y
=If
^ J did^dx
dx,
i,i
dv, dx,
dx+
**a t/^Vj UA,;
S
dx
8Xi
= MV f — 2 dx< N \ \Vv\2 ••=! J n 3 x ;
N /
C
Jn
C
2
\
< - a |v|2rf x + v |Vv| dx ; " \ Jn Jn / we finally obtain a. [ \y\2dx + v [ [Vv|2 rfx > — (* |V • v|2 dx,
Vv
(5.325)
From (5.324), (5.325) it then follows that 12 < lp"l2 - p ( ^ -
(5.326) Since s/ is an automorphism from H onto H (cf. Theorem 5.10), we have (5.327)
If condition (5.313) holds, from (5.326), (5.327) we have (5.328)
with (5.329)
5 Numerical Solution of the Navier-Stokes Equations
305
Relations (5.328), (5.329) prove the convergence properties stated above (and give moreover an estimate of the speed of convergence).
Variants of (5.311), (5.312). In Crouzeix [2] one can find variants of (5.311), (5.312) where a sequence of parameters {pn}n>0 (cyclic, in particular) is used instead of a fixed p. Accelerating method of Tchebycheff type can also be found in the above reference for (5.311), (5.312). Steepest descent and minimal residual procedures for (5.311), (5.312) can be found in Fortin and Glowinski [3] and Fortin and Thomasset [2]. Each of these methods requires N uncoupled Dirichlet problems for a/ — vA to be solved at each iteration. However, these variants of (5.311), (5.312) seem less efficient (at least their discrete analogues) than the conjugate gradient method of Sec. 5.8.7.4.4 which, by the way, is only slightly costlier to implement. 5.8.7.4.4. A conjugate gradient method. From Daniel [1] it follows that one may solve the Stokes problem (5.308) via (5.310) by a conjugate gradient method. Such a conjugate gradient algorithm is defined as follows. Step 0: Initialization p° e L2(Q) arbitrarily given;
(5.330)
au° - vAu° = f - Vp° in Q, u° = g on T,
(5.331)
then solve
and set (5.332)
u°
z° = V
w ° = z.0
Then for n > 0, compute
(5.333)
p"+1,zn+l,wn+* /rom p", z", w" as follows.
Step 1 : Descent /_n ,.,n\ Fn
||z" Wmn) (^wn,
U , ww") )Ln2n)
(s/w",
L (n
p»+1=p»-
- Pn">"
(5.334) (5.335)
Step 2: New descent direction zn+l
=
zn
_
(5.336)
||z" +1 \\h<m 2
Ik" II;L (D) w" +1 = z " + 1 n = n + 1 go to (5.334).
2
'
(5.337) (5.338)
306
VII Least-Squares Solution of Nonlinear Problems
To implement (5.330)-(5.338), it is necessary to know jtfw". From Theorem 5.10 (see Sec. 5.8.7.4.2), sfw" can be obtained by f e (H10(Q))N
aZ" - vAZ" = - Vw" in Q,
(=^> f = 0 on F),
s/wn = V • f.
(5.339) (5.340)
Thus each iteration costs N uncoupled problems for ocl — vA. The strong convergence of {u", p"} to {u, p) can be shown as in Theorem 5.11. Remark 5.24. Due to the H-ellipticity of ssf, it is not necessary to precondition (i.e., to scale) the conjugate gradient algorithm above. Remark 5.25. Algorithm (5.330)-(5.338) is clearly a continuous analogue of algorithm (5.285)-(5.294) in Sec. 5.8.7.3.4. 5.8.7.4.5. Augmented Lagrangian methods. We now consider the continuous analogues of the methods described in Sec. 5.8.7.3.5. Therefore, let r > 0; we note that the Stokes problem (5.222) (and (5.308)) is equivalent to au - vAu - r\(\
• u) + \p = f in Q, V • u = 0 in Q, u = g on F. (5.341)
It is then natural to generalize algorithm (5.311), (5.312) by: p° 6 L 2 (Q) arbitrarily given; n+1
then for n > 0, p" being known, we compute u", p
by
n
au" - vAu" - rV(V • u") = f - \p in Q, u" = g on T, p"+! = p" -
p
v • u",
(5.342)
p > 0.
(5.343) (5.344)
For the convergence of (5.342)-(5.344), we can prove—by a variant of the proof of Theorem 5.11—the following: Theorem 5.12. / / in (5.342)-(5.344), p satisfies 0 < p < l(r + ~\
(5.345)
then, V p° e L2(Q), one has lim {u", p") = {u, p)
in (H^Q))"
x L2(Q) strongly,
where {u, p} is the solution of the Stokes problem (5.222), (5.308) with \npdx — Jn p° dx. Moreover, the convergence is linear. Remark 5.26. The above result can be made more precise by observing that p"+1 -
p
= (/ _ p{rl + S/-1)-
l
)(pn - p).
(5.346)
5 Numerical Solution of the Navier-Stokes Equations
307
Let us prove (5.346); using the notation in the proof of Theorem 5.11, we have p»+i = p " - p\-u",
(5.347)
au" - vAu" - rV(V • u") = - Vp" in Q, u" e (H10(Q.))N.
(5.348)
In fact (5.348) can also be written au" - vAu" = - V(p" - r\ • a") in Q, u" e (if J(Q))N;
(5.349)
since V • 0" £ H, from (5.349) (and from the definition of sf) we have V u" = jaf(pB - r\ • u")
i.e., V • u" = (7 + rsfy Wp".
(5.350)
Combining (5.350) with (5.347), we obtain p» +1 = (/ - p(I + rs4Txs4)vn.
(5.351)
We have
I - p(I + r^Y1^
= (rl + ^/- 1 )" 1 ((r - p)I + s4~x),
(5.352)
and (5.352) yields | | / - p ( / + r j / ) - W | |
ll^" 1 !!).
(5.353)
From (5.351), (5.353) it follows that for the classical choice p = r, we have \\P"+1 - Phnn) < ^ y ^
\\P" - Phnn).
(5.354)
Therefore, if r is large enough and if p = r, the convergence ratio of algorithm (5.342)-(5.344) is of order 1/r. Remark 5.27. The system (5.343) is closely related to the linear elasticity system. Once it is discretized by finite differences or finite elements (as in Sec. 5.8.7.3.5), it can be solved using a Cholesky factorization LL' or LDL', done once and for all. Remark 5.28. Algorithm (5.342)-(5.344) has the drawback of requiring the solution of a system of N partial differential equations coupled (if r > 0) by fV(V-), while this is not so for algorithms of Sees. 5.8.7.4.3 and 5.8.7.4.4. Hence, much more computer storage is required. Remark 5.29. By inspecting (5.354), it seems that one should take p = r and r as large as possible. However, (5.343) and its discrete forms would be ill conditioned if r is too large. In practice, if (5.343) is solved by a direct
308
VII Least-Squares Solution of Nonlinear Problems
method (Gauss, Cholesky), one should take r in the range of 102v to 105v. In such cases, and if p = r, the convergence of (5.342)-(5.344) is extremely fast (about three iterations). Under such conditions, it is not necessary to use a conjugate gradient accelerating scheme.
5.9. Numerical experiments In this section we shall present the results of some numerical experiments obtained using the methods of the above sections. Further numerical results obtained using the same methods may be found in B.G.4P. [1], Le Tallec [2], Periaux [2], Glowinski, Mantel, Periaux, and Pironneau [1], Glowinski, Mantel, Periaux, Perrier, and Pironneau [1]; here we follow Bristeau, Glowinski, Mantel, Periaux, Perrier, and Pironneau [1, Sec. 6]. In Sec. 5.9.1 we shall describe the results related to flows in a channel whose section presents a sudden enlargement due to a step; in Sec. 5.9.2 the numerical results will be related to a flow at Re = 250, around and inside an idealized nozzle at high incidence. 5.9.1. Flows in a channel with a step We consider the solution of the Navier-Stokes equations for the flows of incompressible viscous fluids in the channel with a step of Fig. 5.4. In order to compare our results with those of Hutton [1], we have considered flows at Re = 100 and 191; the computational domain and the boundary condition are also those of Hutton, loc. cit., i.e., u = 0 on the channel walls and Poiseuille flows upstream and downstream. We have used the space discretization of Sec. 5.3 with Vh defined by (5.21) in Sec. 5.3.3.1, i.e., uh (resp., ph) piecewise linear on 2Th (resp., ^ ) ; both triangulations ^ and ^ are shown on Fig. 5.4 on which we have also indicated the number of nodes, finite elements, and nonzero elements in the Cholesky factors of the discrete analogue of — A (resp., I/At — vA) associated with 9~h (resp., 2Ty). As we can see from these numbers, we are really dealing with fairly large matrices; the second of these matrices would have been even larger if we had used (on ^ ) a piecewise quadratic approximation for the velocity. Figure 5.4 also shows the refinement of both triangulations close to the step corner and also behind the step. The steady-state solutions have been obtained via the time integration of the fully discrete Navier-Stokes equations, using those schemes described in Sec. 5.4. We have used, most particularly, the backward implicit scheme of Sec. 5.4.1 and also the alternating-direction schemes of Sees. 5.4.4.1 and 5.4.4.2. The numerical tests have been performed with k = At = 0.4. On Fig. 5.5 we have indicated the stream lines of the computed solutions showing very clearly a recirculation zone whose size increases with Re. If H is the height of the step, we observe that the length of the recirculation zone is approximately 6H at Re = 100 and 87/ at Re = 191, in good agreement with the
/ / / /
l
/
"X
^/
/ /\ \ s\ 7 / \/\ \7/
•J h
N
/
/
\
619 1109 21654
Figure 5.4
Nodes Triangles Cholesky's coefficients
\ \ / / / \ / / / \ \/ / \ \ \ / / / / /I \ / \ / \ \ \/ \ \ \ \ / 7 \ s \/ \ \ / 7 \ \ \'s\\ \/ \\ \\/ \/ \ z/ \ / \ 7 / \ / 7 / / / /\ / /
/ / \ \ \ N/ "/ \ 7
/ / / / / / / \ \ / / / \ / \ \ \ \/\ / / uMu //^
~x \ / \
/• \ / \ / \
\
\ /
\ /
2346 4436 154971
\ /
y \ /
/
/ \ \ \ / / / / \ / / / \ /" NN NN
\ / /
/ \ / / \ \ \ / / / / / / \ N \ / / / / N /
X
~x / / \ \ / / / / x X \ / \ / N o
I
z
VII Least-Squares Solution of Nonlinear Problems
310
Re = 100
Re = 191 Figure 5.5. Stream lines.
results obtained by other authors, and, particularly, Hutton [1]. Finally, in Fig. 5.6, we have shown the pressure distributions corresponding to the two above Reynolds numbers. From these numerical experiments it appears that the alternating-direction schemes are half as costly (in terms of computational time) as the nonsplit schemes; moreover, the incompressibility condition (measured by || V • uh || L2(n)) is also better satisfied with these alternating-direction schemes; as a last remark we would like to mention that the least-squares solution of the nonlinear step in the above alternating-direction methods requires about three conjugate gradient iterations33 (despite this small number of iterations the cost of the nonlinear step is about 10 times larger than the cost of the Stokes step(s)). In conclusion, on the basis of the above examples, we advocate the alternating-direction schemes even for solving the steady-state problems. 5.9.2. Unsteady flow around and inside an idealized nozzle at high incidence
We again follow Periaux [2] and Bristeau, Glowinski, Mantel, Periaux, Perrier, and Pironneau [1] in this subsection. The problem under consideration is the unsteady flow of an incompressible viscous fluid at Re = 250 (see also Glowinski, Mantel, Periaux, Perrier, and Pironneau [1] and Glowinski, Mantel, Periaux, and Pironneau [1] for computations at Re = 750) around and inside an idealized nozzle at high incidence (Re has been computed using the distance between the two walls of the nozzle as the characteristic length). Figures 5.7 and 5.8 show an enlarged view of the nozzle close to the air intake. The computational domain is clearly bounded, and appropriate boundary conditions have been prescribed in the far field. The velocity vector at oo (in the actual problem) is horizontal, and the angle of attack of the nozzle is 40° (which is quite large at that Reynold's number); another feature of our 1
At least for this class of problems.
5 Numerical Solution of the Navier-Stokes Equations
Re = 100
Re = 191 Figure 5.6. Isopressure lines.
Figure 5.7. Details of STh close to the air intake.
311
312
VII Least-Squares Solution of Nonlinear Problems
Figure 5.8. Details of Jh close to the air intake.
problem is that a given flow has been prescribed (via a velocity distribution) on a cross section of the nozzle in order to simulate a suction phenomenon (due to an engine). For the space discretization we have used the same finite element method as in Sec. 5.9.1, and for the time discretization, the CrankNicholson scheme of Sec. 5.4.2, with k = At = 0.1; we have used a triangulation £Th (resp., ^ ) consisting of 1458 (resp., 5832) triangles and 795 (resp., 3049) nodes. Details of 2Th and 9^ are shown on Figs. 5.7 and 5.8. Let us mention that the number of nonzero elements in the Cholesky factors, corresponding to 2Th and - A (resp, $\ and 21/k - vA) is 101 370 (resp, 314 685). Further computational details about the present problem are given in Periaux [2]. As mentioned before, a Crank-Nicholson scheme has been used, taking as the initial value the solution of the corresponding (discrete) Stokes problem (creeping flow). In Fig. 5.9-5.12 we have indicated the distributions at various time steps of the velocity, stream function, pressure, and vorticity, respectively. It is interesting to observe the formation of eddies at the leading edges and their propagation inside and outside the nozzle.
5 Numerical Solution of the Navier-Stokes Equations time step 0 REYNOLDS 250.
time step 20 REYNOLDS 250.
time step 40 REYNOLDS 250.
time step 60 REYNOLDS 250.
313
\
time step 80 REYNOLDS 250.
time step 100 REYNOLDS 250.
Figure 5.9. Unsteady flow for an air intake at high incidence (velocity distribution).
As may be guessed, the computer time required for solving such a large time-dependent nonlinear problem (about 7000 unknowns) on such complicated geometry is important; it actually required several hours of IBM 370-168 to run 100 time steps; however, we have the feeling (on the basis of the comparisons made for the problems of Sec. 5.9.1) that an alternatingdirection scheme would substantially reduce the computer time and could
314
VII Least-Squares Solution of Nonlinear Problems
time step REYNOLDS
0 250.
time step REYNOLDS
20 250.
time step REYNOLDS
40 250.
time step REYNOLDS
60 250.
time step 80 REYNOLDS 250.
time-step REYNOLDS
100 250.
Figure 5.10. Unsteady flow for an air intake at high incidence (stream lines).
provide results of better quality (actually this prediction has been supported by recent numerical experiments at Re = 750 and 1500). As mentioned above, the results of computations at Re = 750 for the same geometry are given in Glowinski, Mantel, Periaux, Perrier, and Pironneau [1] and Glowinski, Mantel, Periaux, and Pironneau [1] (refined triangulations are necessary). Let us also mention that the convergence at each time step of our conjugate gradient algorithm is usually obtained in three to five iterations. To conclude, let us mention that a very important saving in compu-
315
5 Numerical Solution of the Navier-Stokes Equations time step REYNOLDS
0 250.
time step -20 REYNOLDS 250. \
time step REYNOLDS
time step REYNOLDS
\
40 250.
80 250.
time step REYNOLDS
100 250.
Figure 5.11. Unsteady flow for an air intake at high incidence (isopressure lines).
tational time and storage can be obtained from the solution of our very large linear systems by sophisticated iterative methods like those based on incomplete factorizations and, again, conjugate gradient schemes (see, e.g., Manteuffel [1], [2], Glowinski, Mantel, Periaux, Pironneau, and Poirier [1], Glowinski, Mantel, Periaux, Perrier, and Pironneau [1] and the references therein).
VII Least-Squares Solution of Nonlinear Problems
316
time step 0 REYNOLDS 250.
time step 20 REYNOLDS 250.
time step 40 REYNOLDS 250.
time step 60 REYNOLDS 250.
time step 80 REYNOLDS 250.
time step 100 REYNOLDS 250.
Figure 5.12. Unsteady flow for an air intake at high incidence (isovorticity lines).
5 Numerical Solution of the Navier-Stokes Equations
317
5.10. Further comments on Sec. 5 and conclusion
In Sec. 5 we have presented several numerical methods for solving the steady and unsteady Navier-Stokes equations for incompressible viscous flows in the pressure-velocity formulation; the machinery behind these methods may seem a bit complicated, but in fact the resulting methodology is extremely robust for the following reasons: (i) The finite element formulation allows us to handle, in a stable fashion (stable because a variational formulation is used), the Navier-Stokes equations on complicated geometries; moreover, we have used finite element methods leading to error estimates of optimal order for both pressure and velocity. (ii) Fully implicit schemes are used for the time discretization since semiimplicit and explicit schemes resulted in very time-consuming methods. (iii) The resulting nonlinear finite-dimensional problems are solved, via a nonlinear least-squares formulation, by a conjugate gradient method with scaling (i.e., with preconditioning), the scaling operator being a discrete Poisson operator or a discrete Stokes operator, depending on whether or not alternate-direction methods are used. Once again, stability is reinforced, since least-squares formulations act as convexifiers of the original problem (locally at least). Moreover, conjugate gradient algorithms usually have a superlinear convergence (practically quadratic in many cases). (iv) The only linear systems to be solved are related to discrete Poisson and Stokes problems; using the finite element methods of Sees. 5.3 and 5.8, we have been able to reduce the solution of the discrete Stokes problems to the solution of linear subproblems of smaller size associated with symmetric positive-definite matrices, which is also a guaranty of numerical stability (these matrices are obviously factorized once and for all if direct methods of solutions are chosen). From the above properties, the resulting algorithms are very reliable and well suited for simulating separated flows with recirculation in complicated geometries. There is no theoretical difficulties in applying the above techniques to the solution of three-dimensional problems; in this latter case, or in very complicated two-dimensional problems, for solving the very large linear systems involved in our methodology we definitely used sophisticated iterative techniques resulting in a very important saving of computational time and storage. In the present methodology, upwinding has been avoided, since we feared that the corresponding artificial viscosity may "kill" some of the smallscale phenomenon possibly associated with flows at high Reynold's numbers. However, in Appendix 2 one may find several references concerning upwinding and also the description (with some numerical tests) of a new upwinded finite element method of high accuracy which may be useful in computational fluid dynamics.
318
VII Least-Squares Solution of Nonlinear Problems
6. Further Comments on Chapter VII and Conclusion 6.1. Comments
In Sees. 1-5 we have considered the solution of nonlinear problems by least-squares methods; actually linear programming techniques have also been used for solving linear and nonlinear boundary-value problems for partial differential equations. In this direction, we should mention Rosen [1], [2] and the survey paper by Cheung [1] (see also the references in these papers). In our opinion least-squares methods are more natural than linear programming methods since many boundary-value problems for partial differential equations have a natural formulation in Hilbert functional spaces; therefore a least-squares formulation using the corresponding metric of these spaces (or of their dual space) is a most natural approach. Closely related to least-squares method is the method of weighted residuals; a basic reference on this subject is Finlayson [1] (see also Zienkiewicz [1]). In Sec. 3.5 we have described a continuation technique for solving nonlinear problems; this technique has been used in Sec. 3.5.4 for solving the test problem—with turning point—(3.70). For a general theory on the approximation of nonlinear problems with turning points and/or bifurcation points, we refer to Brezzi, Rappaz, and Raviart [l]-[3]. In Sec. 4 we have discussed the least-squares solution of a particular family of compressible flow problems; we refer to Chattot [1] where closely related least-squares methods have also been used for solving nonlinear compressible flow problems. In Sec. 5 we have discussed the numerical solution of the Navier-Stokes equations for incompressible viscous fluids by finite element methods. We have chosen to work with the primitive variables (i.e., pressure and velocity) and with continuous approximations of these variables; we refer to Argyris and Dunne, [1] for a discussion of the finite element solution of the NavierStokes equations using a stream function formulation (see also Reinhart [1] and Bristeau, Glowinski, Periaux, Perrier, and Pironneau [1, Sec. 4.4], and for a survey of finite difference methods, see Roache [1]). For finite element approximation of the Navier-Stokes equations using discontinuous finite element approximations of the velocity and/or the pressure, see Girault and Raviart [1] and the survey paper by Raviart [3]. Concerning alternating-direction and fractional step methods, we would like to mention that the symmetrization process applied in (5.56)-(5.59) and (5.168)—(5.171) is due to Strang [2]; splitting methods for the incompressible Navier-Stokes equations are mathematically analyzed in Temam [1] and Beale and Majda [1]. Basic references concerning alternating-direction methods for solving initial-value problems are34 Beam and Warming [1], [2] • See also the references therein and Godlewsky [1] and Hayes [1], [2].
6 Further Comments on Chapter VII and Conclusion
319
and Warming and Beam [ l ] - [ 3 ] ; these references are in fact strongly motivated by the numerical solution of the compressible Navier-Stokes equations. Returning to the direct solution (discussed in Sec. 5.8.7.2) of the discrete Stokes problem (5.157), i.e.: Find {uh, ph} e Vgh x Hi such that
c
c
c
a. \ uh • \h dx + v \\uh Jn Jn
c
• \\h dx + \Sph • yhdx = f„ • \h dx, Jn Jn n
\-uhqhdx
Vv^e VOh,
= 0,
VqheHlh,
we have seen (in Sec. 5.8.6.2) that (5.157) is in fact equivalent to: Find uh e Vgh such that f uft • \h dx + v n
f Vufc • \\h dx=
Jn
ffc • vfc dx, Jn
Vvȣ VOh,
(6.1)
where VOh = \yheVOh,j\Vgh = k I
yhqh dx = 0,
V qh e Hi
e Vgh, f V • yhqk dx = 0, Jn
\/qhe Hi
(Hi, VOh and Vgh have been defined in Sec. 5.3.3.1). Local vector bases of VOh and Vgh have been found by F. Hecht [1]; they reduce the solution of (5.157) to that of a linear system whose matrix is symmetric, positive definite, and sparse. The solution of this system directly gives uh. As a last comment concerning the Navier-Stokes equations, we would like to mention that the exponential stretching method of Sec. 4.8 can be applied (and has been used, in fact, by several authors) to the solution of the NavierStokes equations on very large space domains (of R2, at least). 6.2. Conclusion In this chapter we have discussed the solution of several nonlinear problems in fluid dynamics by least-squares and conjugate gradient methods. Actually the methods described in this chapter can also be (and have been) applied to the solution of other nonlinear problems. A very important ingredient in the methodology that we have discussed is the choice of a proper norm to use in the least-squares formulation. In fact, after discretization, the corresponding conjugate gradient methods can be classified as conjugate gradient algorithms with scaling (or preconditioning). Related techniques have been used (in a non-least-squares context) to solve linear and nonlinear boundary-value
320
VII Least-Squares Solution of Nonlinear Problems
problems (much simpler than those in this chapter) by, e.g., Bartels and Daniel [1], Axelsson [1], Douglas and Dupont [2], and Concus, Golub, and O'Leary The problems in this chapter are nonlinear, and several steps in the method we have described lead to mathematically open questions. The conjugate gradient method has been chosen, since in some circumstances (see Polak [1] and Daniel [1] for more details) it may achieve a superlinear convergence, like Newton's method, in fact. Compared to Newton's method, the conjugate gradient method with scaling has the advantage of requiring at each iteration the solution of linear systems related to a matrix which is always the same during the iterative process; hence this matrix can be prefactorized (by Cholesky's method, for example).
APPENDIX I
A Brief Introduction to Linear Variational Problems
1. Introduction At various points in the previous chapters we have tried to reduce the solution of nonlinear variational problems to that of sequences of linear variational problems, clearly showing thus the importance of the latter subject. Despite the fact that the theoretical and numerical solutions of linear partial differential equations via linear variational formulations have motivated several excellent books and monographs (let us mention, among others, Mikhlin [1], [2], Lions [2], Necas [1], Aubin [2], Ciarlet [2], Oden and Reddy [1] and Rektoris [1]; see also the references therein1) we felt obliged (for completeness) to give a brief introductory account of this very important subject. For brevity we shall concentrate on elliptic problems (for linear parabolic and hyperbolic problems see, e.g., Lions [3]). In Sec. 2 we shall give the formulation of a standard family of abstract linear variational problems and shall prove existence and uniqueness results. In Sec. 3, which is concerned with the internal approximation of the variational problems of Sec. 2, we shall prove some general approximation results. Finally, in Sec. 4 we shall discuss the variational formulation of some boundary-value problems for linear elliptic partial differential operators (including problems in linear elasticity); some brief comments on the numerical approximation of these problems by finite element methods will complete the discussion.
2. A Family of Linear Variational Problems 2.1. Functional context
We consider: (i) a real Hilbert space V with scalar product (•, •) and associated norm
Ml; (ii) V*, the topological dual space of V; See also Schultz [1] and Prenter [1].
322
App. I A Brief Introduction to Linear Variational Problems
(iii) a bilinear form a: V x V -* R, continuous and K-elliptic (i.e., 3 a > 0 such that a(v, v)> <x\\v\\2, V v e V); a{-, •) is possibly nonsymmetric (i.e., we d o not assume that a(u, v) = a(v, u), V u, v e V); (iv) a linear continuous functional L:V -> M.
2.2. Formulation of the problem The fundamental linear variational problem under consideration is formulated as follows: Find u e V such that a(u, v) = L(v),
VueF.
(2.1) (P)
2.3. Existence and uniqueness results for (P) The following theorem is known as the Lax-Milgram
Theorem.
Theorem 2.1. Under the above hypotheses on V, a(-, •), and L ( ) , problem (P) has a unique solution. PROOF. We first prove the uniqueness and then the existence. (1) Uniqueness. Let ut and u2 be solutions of (P). We then have a(ui, v) = L(v),
VveV,
u^eV,
(2.2)
a(u2, v) = L(v),
VveV,
u2eV.
(2.3)
By substracting, from (2.2), (2.3) we obtain a(u2 - «!, v) = 0,
VueF;
(2.4)
taking v = u2 — u1 in (2.4) and using the F-ellipticity of a(-, •), we obtain ct\\u2 - w!II2 < a(u2 - «!, u2 - «i) = 0, which proves that « t = u2, since a > 0. (2) Existence. From the Riesz representation theorem for Hilbert spaces,2 there exists a unique / e V such that L(v) = (I, v),
V v e V.
(2.5)
2
The Riesz representation theorem states that, V F e F * (i.e., F is linear and continuous from V to R), there exists a unique fe V such that F(v) = (/, v),
V v e V.
For a proof of this fundamental theorem, see, e.g., Yosida [1], Oden [1], or any textbook on functional analysis or Hilbert spaces.
2 A Family of Linear Variational Problems
323
Now suppose that v has been fixed in a(v, w); since the mapping from V to U defined by w -> a(v, w) is linear and continuous, from the above Riesz theorem there exists a unique element of V, denoted by A(v), such that a(v, w) = (A(v), w),
Vt,weK.
(2.6)
Let us prove the linearity of this operator A; on the one hand we have a(A1vl
+ X2v2,w)
= (A(Xlv1
+ X2v2),w),
V Xu l2 e U,
Vvuv2,weV;
(2.7)
on the other hand, from the bilinearity of a(-, •), we have a{kxvi + X2v2, w) = A 1 a(y 1 , vt>) + X2a(v2, w) = ^(A(viX
w) + X2(A(v2),
w)
(2.8)
= (X1A(v1) + X2A{v2),w). From (2.7), (2.8) it follows that A{XiVi + X2v2) = A 1 ^(u 1 ) +
X2A(v2),
3
which shows the linearity of A ; from this linearity property we shall use the notation A(v) = Av. Let us show the continuity of A; from the continuity of a( •, •) there exists M such that \a(v,w)\
< M\\v\\ \\w\\,
Vi), we V,
i.e. (from (2.6)), \(Av,w)\<M\\v\\\\wl
Vv,weV.
(2.9)
Taking w = Av in (2.9), we have \\Av\\ < M\\v\\,
which shows the continuity of A (and also that ||.4|| < M). 4 It follows from (2.5), (2.6) that (P) is equivalent to the linear problem in V, Au = I,
(2.10)
which is, in turn, equivalent to finding u such that u = u — p(Au — I) for some p > 0;
(2.11)
to solve the fixed-point problem (2.11), consider the mapping Wp: V -* V defined by Wp(v) = v-
p(Av - 0;
(2.12)
if vu v2 e V, we have Wp(v2)
3
- WJVl)\\2
=\\v2-
2
Vl\\
- 2pa(v2
- vu v2 -
Vl)
+ P2\\A(v2
-
Vl)\\
2
.
One can easily prove that a(-, •) symmetric (i.e., a(v, w) = a(w, v), V v, w e V) implies the symmetry of A (i.e., (Av, w) = (Aw, v), V v, w e V). 4 W e a l s o h a v e a < \\A\\, since V » E V, w e h a v e a[|p|j 2 < a(v, v) = (Av, v) < \\A\\ \\v\\1.
324
App. I A Brief Introduction to Linear Variational Problems
Hence we have \Wp(v2) - Wp(Vl)\\2 < (1 - 2pa + p2\\A\\2)\[v2 -
vj2;
then Wp is strictly and uniformly contracting if 0 < p < 2a/||^|| 2 . By taking p in this range, we have a unique solution of the fixed-point problem (2.11) which implies the existence of a solution for (P) and completes the proof of the theorem. D
2.4. Remarks Remark 2.1. The Lax-Milgram Theorem (Theorem 2.1) is a particular case of the Lions-Stampacchia Theorems (Theorems 3.1 and 4.1 of Chapter I, Sees. 3 and 4) (obtained by taking K = V and j = 0, respectively). Remark 2.2. Suppose that a(-, •) is a symmetric bilinear form over V x V; we then have: Proposition 2.1. There is equivalence between problem (P) and the following problem: Find u e V such that J(u) < J{v),
\fveV,
(n)
J(v) = ia(v, v) - L(v).
(2.13)
where
PROOF. (1) (P) implies (n). Let u be the solution of (P) and let v e V; we have J(v) = J(u + v — u) = ^a(u + v — u,u + v — u) — L(u + v — u) = \a(u, u) — L(u) + a(u, v — u) — L(v — u) + %a(v — u,v — u) = J(u) + a(u, v - u) - L(v - u) + ja(v - u,v - «). (2.14) Since u is solution of (P) and a(-, •) is F-elliptic, we have a(u, v - u) - Up - u) = 0, a(v-u,
v-u)>0,
V v e V, MveV;
(2.15) (2.16)
from (2.14)-(2.16) it then follows that J(v) - J(u) > 0,
V v e V,
i.e., u is a solution of (n). (2) (n) implies (P). Let u be a solution of (n) and v e V; since u is a solution of (n), we clearly have
2 A Family of Linear Variational Problems
325
Since hm
J(u + tv)-J(u)
= a(u, v) — L(v),
I>0
it follows from (2.17) that a(u, v) - L(v) > 0,
VteF.
(2.18)
Taking — v instead of v in (2.18), we see that a(u, v) - L(v) = 0,
V v e V,
i.e., u is a solution of (P).
•
Remark 2.3. Suppose that V is a complex Hilbert space; then Theorem 2.1 still holds if we suppose that a(-, •) is continuous, sesquilinear (i.e., a(l1v1 + X2v2, w) = Xxa{vx, w) + X2a{v2, w), a(v, A ^ j + X2w2) = l2a(v, w j + X2a{v, w2),
Vf1y2,weF, Vii,w 1 ,w 2 eF,
VA1,A2eC, VAl5 A 2 eC),
and F-elliptic in the following sense 5 : ^a(i;, u) > a||i)|j2,
VteF
with a > 0. Remark 2.4. The hypotheses on V, a(-, •), and L ( ) being those of Sees. 2.1, 2.2, and 2.3, we consider another real Hilbert space H, such that V c H, the injection T 6 from V to H is compact 7 ;
(2.19)
we suppose that H is equipped with the scalar product [•, • ] . Using the Riesz-Schauder theory (see, e.g., Riesz and Nagy [1] and Yosida [1]), we can prove the following generalization of the Lax-Milgram Theorem (Theorem 2.1.). Theorem 2.2. Let XeU and let the hypotheses on V,a(-, •), and L(-)be Sees. 2.1, 2.2, and 2.3. Then the linear variational problem:
5 6
®T. = real part of z e C (*z = x if z = x + iy (x, y e R)). The injection operator T: V -> H is defined by xv = v,
1
those of
V ve V.
That is: (i) T is continuous from F to H. (ii) Let X be a bounded set of V; then the closure X" of X in H is a compact set of H. This implies that if limn^ + co vn = v weakly in V, then limn^ + 0O v» = v strongly in H. See, e.g., Yosida [1] for a general definition of compact operators.
326
App. I A Brief Introduction to Linear Variational Problems
Find u e V such that a(u, v) + l[u, v~] = L(v),
VveV,
(2.20)
has a unique solution if we suppose that the following (eigenvalue) problem: Find w e V, w # 0 such that a(w, v) + A[w, v] = 0,
V v e V,
has no solution.
3. Internal Approximation of Problem (P) In this section we shall study the approximation of (P) from an abstract axiomatic point of view. 3.1. Approximation of V We suppose that we are given a parameter h converging to zero and a family {Vh}h of closed subspaces of V (in practice, the Vh are finite dimensional and h varies over a sequence). We suppose that {Vh}h satisfies the following condition: There exist y rh- ^
cz V such that -f = V, and ->• Vh such that lim \\rhv - v\\ = 0,
V v e V.
(3.1)
Remark 3.1. The terminology "internal approximation" is justified by the fact that Vh c V, V h. 3.2. Approximation of (P) We approximate problem (P) by the following: Find uh e Vh such that a(uh,vh) = L(vh),
VvheVh.
(Ph)
From Theorem 2.1 it then follows that: Theorem 3.1. Problem (Ph) has a unique solution. Remark 3.2. In most applications it will be necessary to replace a(-, •) and L by ah(-, •) and Ln (usually defined, in practical cases, from a(-, •) and L by a numerical integration procedure); on this important matter we refer to Ciarlet [1], [2], [3], Remark 3.3. Suppose that a(-, •) is symmetric; then we can easily prove that the approximate problem (Ph) is equivalent to the minimization problem:
3 Internal Approximation of Problem (P)
327
Find uh e Vh such that J(uh)<J(vh),
\fvheVh,
(nh)
where the functional J is still defined by (2.13). 3.3. Convergence results The convergence results to be proved in a moment follow directly from: Lemma 3.1 (Cea [3]). Let u (resp., nh) be the solution of (P) (resp., (Ph)); we then have K - u|| < ~
Inf \\vh - u\\.
(3.2)
If a(-, •) is symmetric, then instead of (3.2), we have
^f.nf^-,,,1.
,3.3)
PROOF, (i) Proof of (3.2). Let vhe Vh; since u (resp., uh) is a solution of (P) (resp., (Ph)), and since Vh c: V, we have a(u, vh - uh) = L(vh - uh), a(uh, vh - uh) = L(vh - uh), which, by substraction, imply a(uh - u,uh- u) = a(uh - u,vh- u),
V%6F,.
(3.4)
From (3.4) (and from the properties of a(-, •) discussed in Sec. 2.3) it then follows that a||u ft - u | | 2 < a(uh - u , u h - u ) < \\A\\ \\uh - u\\ \\vh - u\\,
V « 4 e Vh,
i.e., a\\uh - u\\ < \\A\\\\vh - M | | ,
\/vheVh,
(3.5)
which, in turn, clearly implies (3.2). (ii) Proof of (3.3). Now suppose that a( •, •) is symmetric; let v e V and u be the solution of (P). From the definition of J (see (2.13)) and from the fact that u solves (P), we then have J(v) = J(u + v — u) = J(u) + a(u, v — u) — L(v — u) + \a(v — u, v — u) = J(u) + $a(v -u,v-u), \/veV. (3.6) Let uh be the solution of (Ph); from Remark 3.3 we have J(vh) > J(uh),
V vh 6 Vh,
i.e., J(vh) - J(u) > J(uh) - J(u),
VvheVh,
which, combined with (3.6), implies that a(uh - u, uh - u) < a(vh - u, vh - u),
V vh e Vh.
(3.7)
328
App. I A Brief Introduction to Linear Variational Problems
From (3.7) and from the properties of a(-, •) we have
a\\uh-u\\2<\\A\\\\vh-u\\2,
V C l eF»;
(3.3) follows clearly from (3.8).
(3.8) •
Lemma 3.1 is quite useful (as shown in, e.g., Ciarlet [1], [2], [3]) for obtaining error estimates; actually it also implies the following: Theorem 3.2. The hypotheses on V, a, and L being those of Sees. 2.1,2.2, and 2.3, and if we suppose that {Vh}h obeys (3.1), we have Urn \\uh -u\\
= 0,
(3.9)
where u (resp., uh) is the solution o/(P) (resp., (Ph)). PROOF. Let us consider e > 0; since V = V, there exists u t e f such that
From (3.10) and Lemma 3.1 it follows that Ik - w|| < — \\rhuc - w|| < — (\\rhue - u,\\ + ||«, - «||) a a < — HMI, - fi,|| + e,
V B > 0.
(3.11)
a Combined with (3.1), relation (3.11) implies 0 < lim sup | K - u\\ < e, V e > 0. The strong convergence result (3.9) easily follows from (3.12).
(3.12) •
3.4. A particular case of internal approximation: the method of Galerkin A popular example of internal approximation is the method of Galerkin described below. In this section we suppose that V is a separable real Hilbert space in the following sense: There exists a countable subset 38 = {Wj}/=" of V such that the subspace V of V generated by 38 is dense in V8 (we can always suppose that the Wj are linearly independent).
8
r = { D 6 F , c = Y&= i A-hWjk with ju > 1}.
(3.13)
3 Internal Approximation of Problem (P)
329
For any integer m > 1 we define 0&m by
and Vm as the subspace of V generated by 3§m; if we denote by nm the projection operator from V to Vm, it follows from the density property stated in (3.13) that l i m \\nmv - v\\ = 0 ,
V c e K
(3.14)
m—*• + oo
The Galerkin approximation (Pm) of problem (P) is then denned as follows: Find um e Vm such that a(um, v) = L(v),
VveVm.
(PJ
Concerning the convergence of {um}m to the solution u of (P), we have the following: Theorem 3.3. We suppose that the hypotheses on V, a, and L are still those of Sees. 2.1,2.2, and 2.3; we also suppose that (3.13) holds. We then have lim \\um - M|| = 0,
(3.15)
m-* + oo
where u (resp., um) is the solution of (P) (resp., (Pm)). PROOF. We can give a direct proof of Theorem 3.3, but in fact it suffices to apply Theorem 3.2 with -T = V, h = 1/m, Vh = Vm, and rh = nm. O
3.5. On the practical solution of the approximate problem (Pfc) A most important step toward the actual solution of problem (P) is the practical solution of (PJ. We suppose that Vh is finite dimensional with Nh = dim Vh.
(3.16)
Let £8h = {wf}fi j be a vector basis of Vh; problem (Ph) is clearly equivalent to: Find uh e Vh such
that a ( u h , W i ) = L(wd,
V ; =1 , . . . , N , .
(3.17) Nh
Since uh e Vh, there exists a unique vector Ah = {XJ}^h1 e U
uh= £ljWj.
such that
(3.18)
330
App. I A Brief Introduction to Linear Variational Problems
Combining (3.17), (3.18), we find that uh is obtained through the solution of the linear system X a(Wj,
Wi)Xj
= L(Wi)
whose unknowns are the Xj,j = \,...
for i = 1 , . . . , Nh
(3.19)
,Nh.
The linear system (3.19) can also be written as follows: A, A, = F,,
(3.20)
where Fh = {L(w;)}f=J1 and where the matrix Ah is defined by Ah = (a(wj, W;))lsijJ-
4. Application to the Solution of Elliptic Problems for Partial Differential Operators 4.1. A trivial example in L2(£l) 4.1.1. Formulation of the problem: existence and uniqueness results In order to illustrate the generalities of Sees. 2 and 3, we have chosen, to begin with, a trivial example. Let Q be a domain 9 of UN (possibly unbounded); we consider the problem (P) associated with the triple {V, a, L} defined as follows: (i) We take V = L 2 (Q); it is a classical result that L2(Q) equipped with the scalar product 10 (v, w) = (* v(x)w(x) dx, Jn and the corresponding norm \112, 9
i.e., an open connected subset of UN. We consider real-valued functions only.
10
V v, w e L2(Q),
(4.1)
\/veL2(Q),
(4.2)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
331
is a Hilbert space (in (4.1), (4.2) we have used the notation dx = dx1 • • • dxN).
(ii) We define a: L2(Q) x L 2 (Q) -> R by a(v, w)=
f ao(x)v(x)w(x) dx,
M v,we L2(Q),
(4.3)
with a0 e LX(D.),
ao(x) > a > 0 a.e. on Q;
(4.4)
a(-, •) is clearly bilinear. From (4.3), (4.4) we have |a(u, w)| < j|a0||Lco(£2)||z;||L2(n)||w||L2(n),
Vu,weL2(a),
(4.5)
and Vt>eL 2 (Q),
a{v,v)>a\\v\\h(Cl),
(4.6)
which imply that a( •, •) is continuous and L2(Q)-elliptic, respectively, (iii) L e t / e L 2 (Q); we finally define L: L 2 (Q) -> U by VoeL2(Q);
L(v) = \ fvdx,
(4.7)
the linear form L is continuous over L 2 (Q) (actually, from the Riesz representation theorem, any linear continuous functional from L 2 (Q) to U has a unique representation of type (4.7)). From Theorem 2.1 (see Sec. 2.3) and from the above properties of L2(Q), a(•, •), and L(•), it follows that the corresponding problem (P), i.e.: Find u e L2(£l) such that f ao(x)u(x)v(x) Jn
dx = f f(x)v(x) Jn
dx,
V U GL2(Q),
(4.8)
has a unique solution. Actually we do not need Theorem 2.1 to see that problem (4.8) has a unique solution which is given by M=
—. a0
(4.9)
Remark 4.1. In the particular case of the bilinear form a(-, •) given by (4.3), (4.4), the operator A introduced in Sec. 2.3 is explicitly given by Av = aov,
VD£L2(Q)
" 1 by
A-1v=—,
VveL2(Q));
332
App. I A Brief Introduction to Linear Variational Problems
furthermore, we have
(and \\A~l\\
1
=
a0
4.1.2. Approximation of problem (4.8) From (4.9) we know the exact solution of problem (4.8); however, for pedagogical purposes, we shall study the approximation of (4.8) by a method of finite element type, giving us an opportunity to apply those general principlesdiscussed in Sec. 3—concerning the approximation of linear variational problems of type (P). For simplicity we suppose that Q is a bounded domain of UN. To approximate problem (4.8) we first introduce a finite family {fi;}f=J i such that Q ; c Q,
Vi=l,...,Nh,
(4.10)!
V i = 1 , . . . , Nh, flj; is a measurable connected subset of Q such
thatj
dx
(4.10)2
>o, 0,
VI
< i,j
i # J,
(4.10)3
Nh
= Q;
(4.10)4
;,
(4.11)
we define h by
where <5, = diameter of Q ; (we recall that diameter Q; =
sup
distance(x, y));
Figure 4.1 illustrates a particular decomposition obeying (4.10) of a domain Q c R2. We now define ** = {w,}f=*i, where V i = 1 , . . . , Nh, W; is the characteristic function of fi;
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
333
Figure 4.1
(we then have (4.13) and w; G LP(Q.% V i = 1 , . . . , Nh, V p such that 1 < p < + oo). We now define Vh as the subspace of L 2 (Q) (and in fact of LP(Q), V p, 1 < p < + oo) generated by 3fih; we then have
Vh = k I »* = J>w,, ",
V i = 1,..., N\,
(4.14)
and also dim Vh = Nh.We observe that if vh e Vh, then vh is piecewise constant. The problem (Ph) corresponding to (4.8) and Vh is denned by: Find uh e Vh such that aouhvh dx = fvh dx, n Ja problem (4.15) has a unique solution.
V vh e Vh;
(4.15)
Concerning the convergence of {w/Jf, to the solution u of (4.8), we have: Proposition 4.1. We suppose that / e L 2 ( Q ) and that a0 satisfies (4.4); we then have lim | K - w||L2(n) = 0, (4.16) where uh (resp., u) is the solution of (4.15) (resp., (4.8)). PROOF. To apply Theorem 3.2 of Sec. 3.3, it suffices to find "V" and rh obeying (3.1). Since L 2 (S1)
C°(fi)
= L2(fi),
we can take "V = C°(Q). We define then rh as follows:
rhveVh,
VveC°(Q
(4.17)!
i
(4.17)2
334
App. I A Brief Introduction to Linear Variational Problems
where, V i = l,...,Nh, prove that 11
P, e Q,; using the uniform continuity of v on fi, we can easily l i m \\rhv - v\\L2ittl *->o
V c e Co(&),
= 0,
which completes the proof of the proposition.
(4.18) •
Suppose that (4.19)
K*=E«IW,.; ;=i
the approximate problem (4.15) is then equivalent to the linear system Z uj doWjWidx = fwtdx, j = 1 Jn Jn
1 < i < Nh.
(4.20)
From (4.10), (4.13) it follows that (4.20) reduces to lit \ a0 dx =
f dx,
Vi=l,...,Nh,
i.e., W - ^ Jn, a 0 dx
Vi = l,...,N f c .
U i =
(4.21)
Remark 4.2. In practice the computation of ut via relation (4.21) would require the computation of the 2Nh integrals below:
f fdx, •JQi
f a0 dx,
1 < i < Nh.
(4.22)
Jat
As mentioned in Sec. 3.2, Remark 3.2, the approximate calculation of the integrals (4.22) would introduce an extra error; also, as mentioned in the above remark, the effect of numerical integration on the solution of elliptic variational problems for partial differential operators is discussed in Ciarlet [1], [2], [3] (see also Strang and Fix [1] and Oden and Reddy [1]). Remark 4.3. If a0 = 1, the solution of (4.8) is clearly u = / and that of the corresponding approximate problem (4.15) is given by: J
1
ax
' We have, in fact, a stronger convergence result since lim \\rhv - v\\Lm(n) = 0.
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
335
it is well k n o w n from integration theory that if uh is obtained from u by (4.23), we have lim \\uh -
M|
LP(n)
= 0,
lim \\uh - M||
V « e LP(O),
= 0,
1 < p < + oo,
VueC°(n)
(4.24) (4.25)
(if u e L°°(Q) we still have (4.24) for all p such that 1 < p < + oo; we also have 12 lim uh = u weakly * in LCO(Q)).
(4.26)
Returning to the case a0 ^ 1, but still obeying (4.4), we can directly prove (using integration theory) that if nh e Vh is given by (4.27) then
lim
= 0,
V/eLp(Q),
l
(4.28)
a0 we can also prove that if fe L°°(Q), then (4.28) still holds for all p e [1, + oo[ and that lim uh = — weakly* in LCO(Q) ( i f a o , / e C o ( Q ) , we have lim
= 0
Mi, —
4.2. Solution of Neumann problems for second-order elliptic partial differential operators In this section we shall discuss the formulation, and the solution via variational methods, of Neumann problems for second-order elliptic partial differential operators; such problems play a very important role in various branches of physics and mechanics.
12
Relation (4.26) means that lim
uhv dx —
uvdx,
see, e.g., Yosida [1] for the properties of the weak* topologies.
336
App. I A Brief Introduction to Linear Variational Problems Figure 4.2
The finite element approximation of these Neumann's problems will be discussed in Sec. 4.5. 4.2.1. The classical formulation Let Q be a (possibly unbounded) domain of RN (we denote by x ( = {x;}f=1) the generic point of UN); we denote by F the boundary dQ. of Q, and we suppose that F is a reasonably smooth manifold with Q locally on one side of F (the last property means that if x0 e F and if we consider a sufficiently small neighborhood °U of x0 in UN, we have the situation of Fig. 4.2). Let A(x) be a linear operator from UN to IR" depending upon x over Q, and let us denote by V the vector {d/cbc;}f=1; we consider the following Neumann problem: Find u such that - V • (K\u) + a0 u = / in Q,
(A°Vu) -n = gonT,
(4.29)
where a 0 , / (resp., g) are given functions defined over Q (resp., F), where n is the outward unit vector normal at F, and where a • b denotes the usual scalar product of IR^, i.e.,
a • b = £ aibh
Va,beR«, a = K}f=u b = {bt}?=1.
i=l
Remark 4.4. The Neumann problem (4.29) is written using the so-called divergence formulation; if S i s a differentiable function of x such that A(x) =
(aij(x))l£ij^N,
from
we have l
CXiOXj
i.e., a quite explicit formulation of the operator v -> V • (SVt;).
i
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
337
Too many engineers and physicists have a tendency to approximate (4.29) (mostly by finite difference methods) using (4.30) to evaluate V • (AVM); actually this approach can be very dangerous since some stability properties of the original problem can be lost during the approximation process, leading to a non-well-posed approximate problem (even if the original problem is well posed); moreover, there is, in general, a conservation principle behind a divergence formulation, and it is of fundamental importance to preserve this conservation property. Another reason for using the divergence formulation is that if A is not differentiable, the divergence formulation may still be meaningful, unlike the product form in (4.30). Actually the variational formulation of problem (4.29)—given in Sec. 4.2.2—provides a quite elegant way to overcome those difficulties related to the possible nondifferentiability of A; also, in some sense, it preserves the above conservation principles. Remark 4.5. If X = I, i.e., if fl
if«' = J,
the Neumann problem (4.29) reduces to — Au + aou = f
in Q,
—= g on
on Y,
(4.31)
where the Laplace operator A is defined by
and where djdn denotes the normal derivative on Y (denned by dv/dn = \v • n). 4.2.2. A variational formulation of the Neumann problem (4.29) Let v be a smooth function denned over Q; multiplying the first relation (4.29) by v and integrating over Q, we obtain -
| V • (K\u)v dx + j aouv dx = \ fv dx. (4.32) •In Jn Jn Let us now consider a vector function x -> V(x), denned over Q and taking its values in UN; we recall (Green-Ostrogradsky formula) that (\-\vdx+ Jn
( v\-Vdx= Jn
f V • (vV) dx = f v\ • n dT, Jn Jr
(4.33)
where dT denotes the superficial measure along Y. Taking V = AVM in (4.33), we obtain
- [ V • (AVH)I; dx = f (K\u) -\vdxJn Jn
f (SVu) • n vdY, Jr
338
App. I A Brief Introduction to Linear Variational Problems
which, combined with (4.32) and with the second relation (4.29), implies
f (K\u) -\vdx+
f a0uvdx= f fv dx + f gv dT;
(4.34)
Jn Jn Jn Jr the function v in (4.32), (4.34) is usually called a test function. We observe that (4.34)—unlike formulation (4.29)—involves only first-order derivatives of the unknown function u. Conversely, it is proved in, e.g., Lions [2], [6] and Necas [1], that if (4.34) holds, V v e y, where 13 y = {v\v e Cl(Q), v has a compact support in Q},
(4.35)
then u is in some sense a solution of the Neumann problem (4.29). At this point, if one wishes to go further, it is necessary to introduce the Sobolev space Hl(Q) defined by 6v
"""
'
(4.36)
the derivatives in (4.36) are taken in the distribution sense, i.e., f ~(j>dx = - [v^-dx, JndXi
V0eS*(Q),
(4.37)
Jn dxi
where S>(Q) = {(/> | <j) e CX(Q), (j> has a compact support in Q}. (4.38) For more details and further properties of the Sobolev spaces, see, e.g., Lions and Magenes [1], Necas [1], Adams [1], and Oden and Reddy [1]. Equipped with the scalar product (», w)HHn) = f (\v • \w + vw) dx Jn and the corresponding norm
(4.39)
the space H J (Q) is an Hilbert space. In the sequel frequent use will be made of the following results (proved in the above references on Sobolev spaces): 13
If n is bounded, then V = Cl(Q). We observe that iffiis bounded (resp., unbounded), //'(D.) contains (resp., does not contain) the functions constant over Q. 14
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
339
Proposition 4.2. We suppose that V is sufficiently smooth (Lipschitz continuous, for example) with Q locally on one side of P. We then have = H\Q),
(4.41)
where S>(0.) = {v\ve CCO(Q), v has a compact support in D.}.15 Let v e ®(Q); we define the trace operator y0 as the linear mapping defined by yov = v\r;
(4.42)
the following proposition can be proved: Proposition 4.3. There exists a constant c(Q) such that V v e ®(fi),
I J ^ l U n < c(fi)||»||w.(Q).
(4.43)
As a corollary to Propositions 4.2 and 4.3, there exists a linear continuous operator from Hl(Q) to L2(T) (the trace operator) whose restriction to ®(Q) coincides with y0; we still use the notation y0 for that trace operator and we have Vcetf^fi),
\\yov\\L2{r) < c(Q)|M| H , (n) ,
(4.44)
with c(Q) as in (4.43). Actually, when no confusion is likely to arise, we shall simply write yov = v. Returning to the Neumann problem (4.29) and to (4.34), we suppose that the following hypotheses concerning A, a0, f, g hold: /eL2(Q), a0 e L°°(Q),
geL\r),
(4.45)
ao(x) > a0 > 0 a.e. on Q,
(4.46)
ai} £ L°°(Q), V 1 < i,j < N; there exists a > 0 such that I
V^ = {^}f=1 e RN
a.e. on Q,
(with |^| = (jtziyy We now define a:Hl{Q) x H 1 ^ ) -• U and L:H\Q) a(v, w) = f (KVv) -\wdx+ L(o) = { fvdx+
f a 0 ^w dx, f 0yo u dF,
-^Uby Vv.we H 1 ^ ) ,
V i; e H^Q),
respectively; a(-, •) (resp., L) is clearly bilinear (resp., linear). ' If n is bounded, we have ®(O) = C°°(n).
(4.47) (4.48) (4.49)
340
App. I A Brief Introduction to Linear Variational Problems
From (4.40), (4.44), (4.45), (4.49), and from the Schwarz inequality in L2(Q) and L2(T), we have
),
(4.50)
which shows the continuity of L. We now define |A(x)| =
Sup
from (4.47) we clearly find that the function x -> | \\x) | belongs to LCO(Q);
(4.51)
we denote by || A||L«,(n) the L^-norm of the above function; we then have, from (4.48) and from the Schwarz inequality, _
_
,
, ,
*1/2
.
\a(v, w)\ <
< Max(|| A|| L « (n) , [|a0||Loo(n))||t;||H.(n)i|w||H1(n),
Vv,we
H\n). (4.52)
Relation (4.52) implies the continuity of a(-, •). Finally, from (4.46), (4.47) we have a{v, v) > Min(a, a 0 )||»|||. (O) ,
Vv e H 1 ^ ) ,
(4.53)
which shows the i? J (Q)-ellipticity of o(-, •)• From the above properties of H1 (fi), a( •, •), and L( •) we can apply Theorem 2.1 of Sec. 2.3 to prove: Proposition 4.4. The linear variational problem: Find u e f / ' f Q ) such that a(u, v) = L(v), V » £ H^Q.),
(4.54)
has a unique solution. Remark 4.6. Suppose that A is symmetric (i.e., a y (x) = a^x), VI < i,j < N, a.e. on Q); this, in turn, implies the symmetry of the bilinear form a{-, •)• From Proposition 2.1 of Sec. 2.4 it then follows that there is equivalence between (4.54) and the minimization problem:
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
341
Find u e H1(Q) such that J(u) < J(v),
V » e H^Q),
(4.55)
where
= - f (K\v) • \v dx + I- f a0v2 dx - f fv dx - [ gv dT. (4. 2 Jn
2 Jn
Jn
Jr
Problem (4.55), (4.56) is a typical problem of the classical calculus of variations. At this step, the natural question that arises is the following: Is the solution of the linear variational problem (4.54) a solution of the Neumann problem (4.29)? To discuss this property, first consider v e ^(Q); since <$(£l) <= H^Q), we have a(u, v) = L(v),
Vi)
or more explicitly, taking into account the fact that y0 v = 0, V v e
f (K\u) -\vdx+
Jn
f aouv dx = f /» dx,
Ja
Jn
V»e ®(O),
(4.57)
i.e., in a distribution sense (see L. Schwartz [1], Necas [1], and Lions [1], [6]) <SV«,Vt;> + (aou, v} = , i>>,
VUG ®(fl),
(4.58)
where < •, • > denotes the duality pairing between 3I'{QL) and 3>{Q). From the properties of the differentiation in @'(Q) (see Schwartz, Lions, and Necas, loc. cit.), we have -
V v e ®(Q),
i.e., in a distribution sense, XV
+
fl0«=/.
(4.59)
A rigorous proof of the fact that the solution u of (4.54) satisfies the boundary condition in (4.29) is definitely beyond the scope of this appendix 16 ; therefore we shall give only a formal proof of this property. Taking v in 2>{Q), from (4.59) we have
-
f V • (X\u)v dx + f aouv dx = f fv dx,
Jn
Jn
Jn
V » e S>(U);
(4.60)
from the Green-Ostrogradsky formula17 (4.33) it then follows that
f (K\u) -\vdx+
Jn
16 17
f aouvdx = f fv dx + f (SV«) • ni; rfr, V i; e
Jn
Jn
Jr
For this proof, see Lions and Necas, loc. cit.
This is precisely the most crucial step, since here we do not justify the fact that (4.33) can be applied in this context.
342
App. I A Brief Introduction to Linear Variational Problems
I.e.,
a(u, v) = f fv dx + f (AVu) • n vdT, Jn
(4.61)
Vȣ
Jr
Comparing (4.61) to (4.54), we obtain vdr = {(K\u)-nv
VteS>(Q);
(4.62)
actually it is proved in the above references that (4.62) implies (AVu) • n = g on F, which is precisely the boundary condition in (4.29). Thus the solution u of the variational problem (4.54) is also a solution of the Neumann problem (4.29); actually we have even more, since the following can be proved: Proposition 4.5. If we suppose that f, g, a0, and A satisfy (4.45)-(4.47), then the Neumann problem (4.29) has a unique solution belonging to Hi(D), which is also the unique solution of the linear variational problem (4.54). Proposition 4.5 is quite important since it suggests the use of (4.54) to solve the Neumann problem (4.29) (for the numerical solution in particular). 4.2.3. Remarks and further comments 4.2.3.1. Some remarks. Remark 4.7. Let us consider N + 1 functions /;, i = 0 , . . . , N, such that /• e L2(fi), V i = 0, ...,N; with {/JfL0 we associate the linear functional L: H\n) -> U defined by
L(v)= hovdx+ £ \ ftp-dx Jn
i=i Jn
clx
i
= f /„ c dx + f f • \v dx, Jn Jn
V v e H^Sl),
(4.63)
where f = {/Jf=1. The linear functional (4.63) is clearly continuous over H1(Q); actually, using the Schwarz inequality, we have A1 N
= 1
1/2
dv IIL* ( O ,
^X~i
N
|| 2 2 L (Q)
+ y
dv
L2(£l)
2
\ 1/2
1=1
1/2
V»e
(4.64)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
343
In fact it is shown in, e.g., Lions [2] that any continuous linear functional over H1(Q) has a (nonunique) representation of type (4.63). We now consider the linear variational problem: Find it e H1(Q) such that a(u, v) = L(v),
V v e H^Q),
(4.65)
where a(-, •) (resp., L(-)) is denned by (4.48) (resp., (4.63)); if (4.46), (4.47) still hold, problem (4.65) has a unique solution. Taking v e S?(Q) in (4.65), we should prove, as in Sec. 4.2.2, that u is a solution (in the sense of distributions) of the linear partial differential equation (of elliptic type) - V • (XVu) + aou = /„ - V • f;
(4.66)
concerning the boundary conditions, we should prove (at least in a formal way) that u satisfies (K\u)
n = f n on T.
(4.67) 2
iV
We observe that if f is the more general function of (L (Q)) , then f • n is meaningless on F ; however, it can be proved that if fe(L 2 (Q)f,
V-feL2(Q),
then a mathematical sense can be given to f • n on F and to the boundary condition (4.67). Remark 4.8. Some application lead us to consider linear functional L(-) like Uv) = f fvdx + \gvdT + \hv dy,
(4.68)
where y is a measurable part of a (JV - 1) manifold contained in Q. We suppose that / , g, and h belong to L 2 (Q), L 2 (F), and L2(y), respectively, which implies that L(-) is continuous over /f X(Q); therefore if we suppose that a(-, •) and L(-) are defined by (4.48) and (4.68), respectively, and that (4.46), (4.47) still hold, then the variational problem: Find u e H1(Q) such that a(u, v) = L(v),
V v e H\Q),
(4.69)
has a unique solution; this solution is also the unique solution, in H1(Q), of - \ • (AVM) + aou = f + T(h) (in the sense of distributions), (K\u) • n = g on F, (4.70) where the distribution T(h) is defined by , v} = [hvdy, Jy
Vt)6
344
App. I A Brief Introduction to Linear Variational Problems
We illustrate the above considerations by the particular case where N = 1, Q = ]0, 1[ and where L(v) = f fv dx + AD(0) + nv(\) + vv(x0),
V » e H\0, 1),
(4.71)
•Jo
a(v, w) =
C1
Jo
Ol(x)
f1
dv dw
— ^-dx+\aovwdx, dx dx Jo
Vc,w£ H\0, 1),
(4.72)
with / e L 2 ( 0 , 1); X, fi,veU;
x o e ] 0 , 1[; a, e L°°(0, 1), ax(x) > a > 0
a.e. o n ] 0 , 1[; a0 e L°°(0, 1), ao(x) > a 0 > 0 a.e. on ]0, 1[.
(4.73)
Since H\0, 1) c C°[0, 1] and the injection from HHO, 1) to C°[0, 1] is continuous 18 (with C°[0, 1] equipped with the L°°(0, l)-norm), (4.71) makes sense (we have to keep in mind that if D. <= UN with N > 2, then H \Q) <£ From the hypotheses (4.73), the linear functional (4.71) (resp., the bilinear functional (4.72)) is continuous (resp., continuous and H\0, l)-elliptic); the variational problem: Find u e H\0, 1) such that a(u, v) = L(v),
V « e H1®, 1),
(4.74)
has a unique solution which is also the unique solution in //*((), 1) of the one-dimensional Neumann problem a0u = f + vdXo in ]0, 1[,
where dxo denotes the Dirac measure at x0. Remark 4.9. Actually the hypothesis ao(x) > a0 > 0 a.e. on Q (cf. (4.46)) is quite strong and can be replaced by the weaker one, ao(x) > 0 a.e. on Q,
ao(x) dx > 0;
' In fact the injection from H'(0, 1) to C°[0, 1] is compact. We can also prove that ll»llco[o.i]f= Max \v(x)\) < V5|[P||H.(O.1,, \
*6[0, 1]
/
Vt6H'(0, 1).
(4.76)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
345
on the other hand, in view of applications (in fluid mechanics, for example) it is quite important to discuss the solution of the Neumann problem (4.29) in the important case where a0 = 0. Before discussing (in Sec. 4.2.3.2) the solution of the Neumann problem (4.29) under the above hypotheses, we shall prove a quite useful (and classical) lemma: Lemma 4.1. Let ilbe a bounded domain of UN, and let us consider the bilinear form a:H\Q) x H\Q) -> U, where a(v, w) = ax{v, w) + aQ(v, w),
V v, w e H\Q);
(4.77)
we suppose that (v, w)=
f (A\v) • Vw dx,
V»,we
H\D.),
•JSl
where A satisfies (4.47), and also ao( •, •) is bilinear, continuous, and positive (i.e., ao(v, v) > 0,
semidefinite
V c e H ^ Q ) ) , and such that
ao(v, v) = 0, v = constant, implies v = 0. We then have the following
(4.79)
property:
the bilinear form a(•, •) is Hl{D)-elliptic.
(4.80)
PROOF. The proof of Lemma 4.1 is quite classical and follows directly from the compactness property below (known as Rellich's Theorem). Since £1 is bounded, we have: the injection from H^Q.) to L2(Q) is compact,
(4.81)
which implies, as a corollary, / / lim vn = v weakly in H^Q),
then lim ||»B - v\\L2(n) = 0.
(4.82)
The / / 1(Q)-eIlipticity of a(-, •) indicates the existence of y > 0 such that a{v, v) > y\\v\\2HHn),
V v e H^Q).
(4.83)
Suppose that (4.83) is not true; there is equivalence between the fact that (4.83) does not hold and the existence of a sequence {vn}n >0 in //'(D.) such that lkllff.
V n,
lim a(vn, vn) = 0.
(4.84)
n-y+ oo
Since—from (4.84)—the sequence {un}n>0 is bounded in the Hilbert space Hl(il), we can extract from {vn}n20 a subsequence—still denoted by {yn}n>0—such that lim vn = v* weakly in H\Q).19
(4.85)
19 This follows from the fact, proved in, e.g., Yosida [1], that the closed, convex, bounded sets of the Hilbert spaces are weakly compact (here {fn}n>0 i s contained in the closed ball of H'(Ci), whose center is 0 and radius 1).
346
App. I A Brief Introduction to Linear Variational Problems
We observe (see part (2) of the proof of Theorem 5.2 in Chapter I, Sec. 5.4) that (4.84), (4.85), combined with the continuity of a(-, •) and its positive-semidefinite property, imply 0 < a(v*, v*) < lim inf a(vn, vn) = lim a(vn, vn) = 0, o-> +
n~* + oo
GO
i.e., a(v*, v*) = 0.
(4.86)
If we make (4.86) explicit, from (4.77), (4.78) we have
(* (K\v*) • \v* dx + ao(v*, v*) = 0.
(4.87)
Combining (4.47), (4.79), and (4.87), we obtain Vn* = 0 a.e. on Q. (
(4.88)
ao(v*, v*) = 0,
(4.89)
and (4.79), (4.88), (4.89), in turn, imply that v* = 0.
(4.90)
We observe that (4.47) and the second relation (4.84) (resp., (4.82), (4.85), (4.90)) imply lim \vn = 0 strongly in (L2(Q))N
(4.91)
lim vn = 0 stongly in L2{Q)).
(4.92)
(resp.,
n-* + ao
From (4.91), (4.92) we have lim v„ = 0 strongly in H1(Q.), n-> + oo
i.e., lim IIBJIH.,,, = 0.
(4.93)
« - • + GO
Actually there is a contradiction between (4.93) and the first relation (4.84); therefore (4.84) does not hold, or equivalently, (4.83) holds. • Applications of Lemma 4.1 are given in the following sections. 4.2.3.2. Some applications of Lemma 4.1. We now apply Lemma 4.1 to the solution of the Neumann problem (4.29) if either a0 satisfies (4.76) or a0 = 0 on Q. More precisely, we have the following propositions: Proposition 4.6. We consider the Neumann problem (4.29) with f, g, A still obeying (4.45), (4.47), respectively; if we suppose that fi is bounded and that a0 e LCO(Q), ao(x) > 0 a.e. on Q,!,
Jn
a o(y, (x) dx > 0,
(4.94)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
347
then (4.29) has a unique solution in H1(Q), which is also the unique solution of the variational problem: Find u e H^Q) such that
I ( X V M ) • \ v dx + I aouv dx = f fv dx + j gv dT, Jn Jn Jn Jr
V H E H\Q). (4.95) 1
PROOF. It suffices to prove that the bilinear form occurring in (4.95) is H (Q)-elliptic. This follows directly from Lemma 4.1 and from the fact that if v = constant = C, then ao(x) dx > 0 •Jn
and
•'n
ao(x)v2 dx = C2
•'n
ao(x) dx = 0
imply C = 0, i.e., v = 0.
•
Proposition 4.7. We consider the Neumann problem (4.29) with Q bounded and A still obeying (4.47); if we suppose that a0 = 0, then (4.29) has a unique solution u in H^nyU20 if and only if
f fdx+ LLdT = 0; ;
Jn
(4.96)
JT
u is also the unique solution in H1(Q)/U of the variational problem: Find u e f / ' P such that
f (SVu) - \ v d x = f fv dx + f gv dT, Jn Jn Jr
Vce
H\£l).
(4.97)
PROOF. For clarity we divide the proof into several steps. Step 1. Suppose that a0 = 0; if u is a solution of (4.29) and if C is a constant, it is clear from V(« + C) = \u that u + C is also a solution of (4.29). If u is a solution of (4.29), we can show, as in Sec. 4.2.2, that (4.97) holds; taking v = 1 in (4.97), we obtain (4.96). Step 2. Consider the bilinear form over Hl(ii) x H 1(Q.) defined by
r
Ir
a(v, w)= \ \v • Vw dx + (\ v dx
1
\/r
\ w dx ,
Vt,w6 H^O);
This means that u is determined in Hl(Q) only to within an arbitrary constant.
(4.98)
348
App. I A Brief Introduction to Linear Variational Problems
the bilinear form a(-, •) is clearly continuous, and from Lemma 4.1, it is if'(fi)-elliptic (it suffices to observe that if v = constant = C, then 0= [ \ vdx] = C 2 (meas(Q))2 => C = v = 0). Vn / From these properties, v -> {a(v, v))112 defines, over if l(Q), a norm equivalent to the usual H^nynorm defined by (4.40). Step 3. We now consider the space
Fj = ivlveH^O), Vi being the kernel of the linear continuous functional i; ->
v(x) dx
i
is a closed subspace of H (il). If we suppose that Hl(Q) has been equipped with the scalar product defined by a( •, •) (see (4.98)), it follows from Step 2, and from the definition of Vlt that over Vlt v 1/2
defines a norm equivalent to the H '(Sll-norm (4.49); henceforth we shall endow Vx with the following scalar product:
7'
• Vw dx.
•In
From these properties of Vu and from (4.47), the variational problem: Find ueV^ such that f (KVu) • \v dx = [fvdx+
[gvdT,
VueFj,
(4.99)
has a unique solution. Step 4. Returning to Hl (Q) equipped with its usual product, let us introduce Vo a H1 (Q) defined by Vo = {v\v = constant over
fi};
if v e VQ , we have r
t
0 = \ \v-Vcdx+
r
\ vc dx = = c \I ;v dx,
which shows that Kg = Vt. We then have
if \Q) =V0® Vu J
and for any v e if (f2), we have a unique decomposition t> = v0 + vu
vteVi,
V i = 0, 1.
VceR,
(4.100)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
349
Step 5. From (4.96), (4.99) it follows that we also have
f (AVu) • V(v + c)dx=
f f(v + c)dx + \g(v + c)dT, V v e Vu
V c e R, u € Vx. (4.101)
From the results of Step 4, relation (4.101) implies that u is a solution of (4.97) (but the only one belonging to F J ; actually if we consider a second solution of (4.97), say u*, we clearly have (from (4.47)) a f \V(u* - u)\2 dx < f (XV(«* - «)) • V(«* - u) dx = 0.
(4.102)
From (4.102) it follows that u* — u = const; this completes the proof of the proposition.
• Remark 4.10. In many cases where a0 = 0 in (4.29), one is more interested in Su than in u itself [this is the case, for example, in fluid mechanics (resp., electrostatics), where u would be a velocity potential (resp., an electrical potential) and Vu (resp., — Vu) the corresponding velocity (resp., electrical field)]; in such cases, the fact that u is determined only to within an arbitrary constant does not matter, since V(« + c) = Vu, V c e R .
4.3. Solution of Dirichlet problems for second-order elliptic partial differential operators We shall now discuss the formulation and the solution via variational methods of Dirichlet problems for linear second-order elliptic partial differential operators. The finite element approximation of these problems will be discussed in Sec. 4.5. 4.3.1. The classical formulation. With Q, A, a0, f, g, and the notation as in Sec. 4.2.1, we consider the following Dirichlet problem: - V • (XVu) + V • (PM) + aou = / in Q,
u = g on F,
(4.103)
where P is a given vector function denned over Q and taking its values in UN. Remark 4.4 of Sec. 4.2.1 still holds for the Dirichlet problem (4.103).
Remark 4.11. If X = I, a0 = 0, p = 0, the Dirichlet problem (4.103) reduces to -Au = / in Q,
u = g on T,
which is the classical Dirichlet problem for the Laplace operator A.
(4.104)
350
App. I A Brief Introduction to Linear Variational Problems
4.3.2. A variational formulation of the Dirichlet problem (4.103) Let v e 9{Q) (where ®(Q) is still denned by (4.38)); we then have c = 0onr.
(4.105)
Multiplying the first relation (4.103) by v, we obtain (still using the GreenOstrogradsky formula, and taking (4.105) into account)
[ (SVu) • \v dx - tufl-Vvdx+
Jn
.
J(i
[aouvdx = [ fv dx,
Jn
Jn
Vi>eS>(Q). (4.106)
Conversely it can be proved that if (4.106) holds, then u satisfies the secondorder partial differential equation in (4.103) (at least in a distribution sense). Let us now introduce the Sobolev space HQ(Q) defined by Hl0(Q) = S>(Q)H'(n);
(4.107)
if F = dQ is sufficiently smooth, we also have Hj(fi) = {v\veHHQ), yo^ = 0},
(4.108)
where y0 is the trace operator introduced in Sec. 4.2.2. From (4.107), (4.108), HQ(Q) is a closed subspace of HJ(Q). An important property of HQ(Q) is the following: Suppose that Q. is bounded in at least one direction of UN; then \l/2
\Vv\2 dx) Ja
(4.109)
)
defines a norm over Hj(Q) equivalent to the H1(Q)-norm. Property (4.109) holds, for example, for Q = ]a, /?[ x R^" 1
with a, jS e U, a < )3,
but does not hold for
Returning to the Dirichlet problem (4.103), and to (4.106), we suppose that the following hypotheses on A, a0, P, / , and g hold: feL2(Q),
lgeH\Q)
a0 e L^fi),
such that g = yog,
ao(x) > a0 > 0 a.e. on Q,
(4.110) (4.111)
X satisfies (4.47),
(4.112)
p e (L (Q)) , V • p = 0 (in the distribution sense).
(4.113)
OO
/V
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
We now define a: H\Q) a(v,w)=
x H\SX) -> U and L: Hl(ii)
\ (K\v) • Vw dx - \v$-\wdx+ Jn Jn
351
-^Uby
\ aQvw dx, Jfi
Vu.wer/'P, (4.114)
V o e / f 1 (Q),
L(v) = f /t> dx,
(4.115)
respectively; a(-, •) (resp., L) is clearly bilinear continuous (resp., linear continuous). Before discussing the variational formulation of the Dirichlet problem (4.103), we shall prove the following useful lemma: Lemma 4.2. Suppose that p satisfies (4.113); we then have
| \ p - \wdx = - f wP- Vvdx, Jn Jn (i.e., the bilinear form {v, w} -+
Vu,wG/fJ(Q)
(4.116)
up • Vw dx Jn
is skew symmetric over HQ(Q,)
X
Hl(Q)).
PROOF. Let v, w e 3){Q); we have I up • Vw dx = I P • \(vw) dxJn Jn Since vw e @(£l) and V • p = 0, we also have
f p • \(vw)
I wp • \v dx. •'n
dx =
(4.117)
(4.118)
(where < •, • > denotes the duality between £)'(Q.) and From (4.117), (4.118), we then have -\wdx=
-
\ wfl-\vdx, Jn
V»,tf6 9(fi).
From the density of @(£l) in tf J(J2), (4.119) implies (4.116).
(4.119) •
Lemma 4.2 implies: Proposition 4.8. Suppose that (4.Ill)—(4.113) hold; then the bilinear a{-, •) defined by (4.114) fs Hl(Q.)-elliptic.
form
PROOF. From (4.111), (4.112), (4.114) we have a(v, v) > Min(a, a o )|M|£i (n) -
(" up • Vu rfx, VUG Hj(fl);
(4.120)
352
App. I A Brief Introduction to Linear Variational Problems
since p satisfies (4.113), from Lemma 4.2 we have yp • Vv dx = 0,
V v e H&Q).
(4.121)
[2
Combining (4.120) and (4.121), we have a(v, v) > Min(a, a o )||y|||i ( jj),
V » e Hj(fl),
i.e., the HJ(Q)-ellipticity of a{ •, •).
•
Using the above results we n o w prove: Proposition 4.9. Suppose that (4.110)—(4.115) hold; then the variational problem: Find u e H1(Q) such that you — g and a(u,v) = L(v),
Vcer/JfQ)
(4.122) 1
has a unique solution. This solution is also the unique solution in H (D.) of the Dirichlet problem (4.103). PROOF. (1) Uniqueness. Suppose that (4.122) has two solutions ut and u2; we then have a(u!, v) = L(v),
V v e HQ(Q),
a(u2,v) = L(v),
Vt)eHj(Q),
and by substraction, a(u2 -u1,v)
= 0,
Vv e Hj(Q).
(4.123)
1
Since uu u2 e H (Q) with youl = you2 (=g), we have u2 — ul eH1^),
yo(u2 — uy) = 0,
e
i.e., « 2 — "i HQ(Q). Taking v = u2 — ut in (4.123) and using Proposition 4.8, we have 0 < Min(a, a o )||« 2 - "IIIHUD) ^ 0, i.e., u2 = «!•
(2) Existence. We have (from (4.110)) u = yo0> where g e H1(Q); this leads to the introduction of u e HQ(Q) such that u = u-g
{&u = u + g).
(4.124)
There is clearly equivalence between (4.122) and the linear variational problem in Hl(il) below: Find u e Hj(fi) such that a(u, v) = L(v) - a(g, v),
V v e H^(£i).
(4.125)
From Proposition 4.8, a{-, •) is bilinear, continuous over Hj(Q), and i/J(fl)-elliptic; moreover, the linear functional v -* L(v) - a(g, v)
(4.126)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
353
is clearly continuous over Ho(ii). From the properties of HQ(Q), a(-, •), and of the linear functional (4.126), we can apply Theorem 2.1 of Sec. 2.3 to prove that (4.125) has a unique solution in HQ(Q); this, in turn, implies (taking part (1) into account) that (4.122) has a unique solution as well. (3) The Solution u of (4.122) Satisfies (4.103) and Conversely. Taking v e &(Q) in (4.122), we find that u satisfies (4.103) in the sense of distributions. Conversely, if u e H1(Q.), with you = g, satisfies (4.103), we can easily prove that a(u, v) = L(v),
V D £ ®(Q),
(4.127)
and using the density of &(Q) in tfj(fi), we find that (4.127) also holds for all the v in
Hj(fl).
•
Remark 4.12. Suppose that A is symmetric and that P = 0; this implies the symmetry of the bilinear form a(-, •). There is then equivalence between (4.122) and the minimization problem: Find u e Hg such that J(u)<J(v),
~iveHg,
(4.128)
where
J(V) = I f (K\v) -\vdx + l f a0v2 dx - f fv dx 2 Jn
2 Jn
Jn
(4.129)
and Hg = {v\veH1(Q),yov
= g}.
(4.130)
4.3.3. Further remarks and comments Remarks 4.7 and 4.8 of Sec. 4.2.3.1 still hold for the Dirichlet problem (4.103) in that we can replace L defined by (4.115) by more complicated linear continuous functionals like the one in (4.63), or
L(v) = f fv dx + [hv dy Jn
Jy
with h and y as in (4.68). Concerning the case where a0 = 0 in (4.103), (4.114), we have the following: Proposition 4.10. Suppose that Q is bounded in at least one direction of UN; also suppose that a0 = 0, the hypotheses on A, P, / , g remaining the same. Then the variational problem (4.122) still has a unique solution which is also the unique solution in H J (Q) of the Dirichlet problem (4.103). PROOF. It suffices to prove that the bilinear functional a: H^Sl) x H^il) -> R defined by
a(v, w) =
(A\v) -\wdx - \ vfl • \w dx
354
App. I A Brief Introduction to Linear Variational Problems
is Ho(£l)- elliptic. This follows from (4.112) (and (4.47)) and from Lemma 4.2 which implies that a(v, (v, v)>a\ v) > a
f\Vv\2 dx,
V v e tf£(Q),
and from (4.109) which implies that v -> (J n | Vu | 2 dx)1/2 is a norm over HJ(Q) equivalent to the H1(Q)-norm. D
4.3.4. Some comments on mixed boundary-value problems (Neumann-Dirichlet problems) We briefly discuss the solution (via variational methods) of elliptic problems for linear second-order partial differential operators combining the boundary conditions of Sec. 4.2 (Neumann's boundary conditions) and Sees. 4.3.1., 4.3.2, and 4.3.3 (Dirichlet's boundary conditions). With fi as in the above sections, we suppose that F (= d£l) is the union of F o , Fj such that F o u Fi = F,
ronr, =0;
(4.131)
such a situation is described in Fig. 4.1 of Chapter II, Sec. 4.1. We now consider the mixed boundary-value problem — V • (AVu) + aou = / i n D.,
u = g0 o n F 0 ,
(AVM) • n = gl on r 1 ( (4.132)
where A, a o , / a r e as in Sec. 4.3.1 and where g0, gx are given functions defined over F o , Tu respectively. Taking a test function v "sufficiently smooth" and such that v\ro=0,
(4.133)
we obtain (still using the Green-Ostrogradsky formula (4.33)) that any solution of (4.132) satisfies
f (SVw) • \v dx + f aouv dx = f fv dx + f grv dY.
Jn
Jn
Ja
Jr t
(4.134)
To solve (4.132) by variational methods, we first introduce a(-, •) and L(-) defined by a(v,w)=
I (KVv) • Vw dx + \ aovw dx, Jn Jn L(v) = [ fv dx + f
giv
dT,
V v, w e H\Q.), VceH'p.
(4.135) (4.136)
We suppose that a0 and A still satisfy (4.46), (4.47) and that /eL2(Q),
g.eViT,).
(4.137)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
355
These hypotheses on A,ao,f,g1 imply that a(-, •) is bilinear continuous over H*(Q) x H^Q) and H1(Q)-elliptic and that L is a linear continuous functional
over H^Q). We now introduce (motivated by (4.133)) the following subspace Vo of
H\ny. Vo = {v|v e H^Q), yov = O a.e. on F o } ;
(4.138)
1
actually Vo is a closed subspace of H (Q). Using a variant of the proof of Proposition 4.9 in Sec. 4.3.2, we should prove the following: Proposition 4.11. Suppose that there exists g0 e H1(Q) such that 0o = yO<7olro;
(4-139)
also suppose that the above hypotheses on A, a0, f, g1 hold. Then the variational problem: Find u e H 1 (Q) such that you = g0 on F o and f (K\u) • \v dx + \aouvdx= Jn
Jn
\ fv dx + j gxvdT, J
"
JFl
V v e Vo (4.140)
has a unique solution, which is also the unique solution in H\Q) of the mixed boundary-value problem (4.132). For a proof see, e.g., Necas [1]. Most remarks of Sees. 4.2.3 and 4.3.3 still hold for (4.132), (4.140); in particular, if D. is bounded and if | r o dT > 0, then we can suppose that a0 = 0 [this follows from the fact that if Q, is bounded, then
\\v\2dxV'2 defines a norm over Vo which is equivalent to the H^Q^norm (Lemma 4.1 can be used to prove this equivalence property)].
4.4. Solution of second-order elliptic problems with Fourier boundary conditions 4.4.1. Synopsis In this section we shall discuss the solution—via variational methods—of the so-called Fourier problem for linear second-order elliptic partial differential operators; our interest in this problem is twofold: (i) The Fourier problem occurs in the modelling of several heat-transfer phenomana.
356
App. I A Brief Introduction to Linear Variational Problems
(ii) The Neumann and Dirichlet problems discussed in Sees. 4.2 and 4.3, respectively, can be considered, in fact, to be particular cases21 of the Fourier problem; this claim will be justified in the sequel. 4.4.2. Classical formulations of the Fourier problem With Q, A, a0, f, g as in Sec. 4.2, we consider the following boundary-value problem: Find u such that - V • (AVM) + a0u=
f in Q,
(X\u)
• n + ku = g onT,
(4.141)
where k is a given function denned over F. 4.4.3. A variational formulation of the Fourier problem (4.141) Let v be a smooth function defined over Q; multiplying the first relation (4.141) by v and integrating over Q, we obtain (using the Green-Ostrogradsky formula (4.33)) f
=
(A\u)-\vdx
+
f
f
C =
aouvdx= fvdx+ (AVw) • nv dT; (4.142) Jn Jn Jn Jr combining (4.142) and the boundary condition in (4.141), we finally obtain | (X\u) • \v dx + \ a0 uv dx + \ kuv dT = \ fv dx + \ gv dT. (4.143) Jn Jn Jr Jn Jr Conversely, it can be proved that if (4.143) holds, V u e f , where V is still defined by (4.35), then u is in some sense a solution of the Fourier problem (4.141). We now define a: H\Q) x Hl(Q) -» R and L: H^O) -> U by
f a(v, w) =
=
f
[(AVu) • Vw + aoi;w] dx + Jn
kvw dT,
Vtt,we H1(Q),
Jr (4.144)
L(v) = f fv dx + f gyovdT,
V v e H!(Q),
(4.145)
Jn Jr respectively. We suppose that the following hypotheses concerning X, a0, k, f, g hold: /eL 2 (Q), geL2(T), a0 e L°°(fi), ao(x) > a0 > 0 a.e. on Q, CO
(4.146) (4.147)
k G L (F), k(x) > 0 a.e. on F,
(4.148)
X satisfies (4.47).
(4.149)
Similarly for the mixed boundary-value problem of Sec. 4.3.4.
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
357
From the above hypotheses and from the continuity of the trace operator y0: H1(Q) -> L2{T) (see Sec. 4.2.2), we find that a(-, •) is bilinear continuous over H1(Q) x H1(Q) and /^(fij-elliptic, and that L(-) is linear continuous; we can therefore apply Theorem 2.1 of Sec. 2.3 to prove: Proposition 4.12. / / the above hypotheses on A, a0, k, f, g hold, the linear variational problem: Find UE H\Q) such that a(u, v) = L(v),
V v e H\Q)
(4.150)
has a unique solution; this solution is also the unique solution in H *(Q) of the Fourier problem (4.141). Remark 4.13. Suppose that A is symmetric; this, in turn implies the symmetry of the bilinear form a( •, • )• From Proposition 2.1 of Sec. 2.4 it then follows that (4.150) is equivalent to the minimization problem: Find u e H^Q) such that J(u)<J(v),
VveH^n),
(4.151)
where
J(v) = i f l(K\v) • \v + a0v2^ dx + \ f kv2 dY - f fv dx - [gv dT. (4.152) Remark 4.14. Using Lemma 4.1, we should prove that Proposition 4.6 of Sec. 4.2.3.2 still holds for problem (4.141), (4.150) in that if Q is bounded and if a0 e LCO(Q), ao(x) > 0 a.e. on Q,
ao(x) dx > 0, Jn then problem (4.141) has a unique solution in H^Q) which is also the unique solution of the variational problem (4.150). Concerning the case a0 = 0 on Q, we shall prove the following: Proposition 4.13. We suppose that Q is bounded and that a0 = 0 on Q; we also suppose that k satisfies fcGL°°(r),
k(x) > 0 a.e. on T,
f k(x) dY > 0.
(4.153)
Then, the hypotheses onf, g, A being (4.146), (4.149), respectively, the variational problem: Find u e Hl(Q) such that
f (SVu) • \v dx + [kuvdY=
Jn
Jr
[ fv dx + f gv dY,
Jn
Jr
V v e H\n), (4.154)
358
App. I A Brief Introduction to Linear Variational Problems
has a unique solution which is also the unique solution in Hl(Q) of the boundaryvalue problem - V • (XVu) = / in Q,
(X VM) • n + ku = g on T.
(4.155)
PROOF. It suffices to prove that the bilinear form occurring in (4.154) is H^^-elliptic. This follows directly from Lemma 4.1 (see Sec. 4.2.3.2) and from the fact that if v = constant = C, then
h
x)dT>Q
and f k(x)v2 dT = C2 f k(x) dT = 0 Jr Jr imply that C = 0, i.e., v = 0.
•
4.4.4. T/je Neumann and Dirichlet problems as particular cases of the Fourier problem We would like to justify the claims of Sec. 4.4.1 concerning the relationships between the Neumann and Dirichlet problems and the Fourier problem. Obtaining the Neumann problem (4.29) from the Fourier problem is quite trivial since it suffices to take k = 0 on F in (4.141) and (4.144); obtaining the Dirichlet problem from the Fourier problem is less simple. First consider the variational formulation of the standard nonhomogeneous Dirichlet problem: Find u e H1(Q) such that you = g, and
1'
[(AVu) -\v + a0 uif] dx = L(v),
V v e HJ(Q),
(4.156)
Jn
where g = yog, g e H^Q.), and where L is linear continuous from H^Q) to U; taking L linear continuous from Hl{Q) to U in (4.156) is not restrictive since any functional linear continuous over Hl(Q) can be considered as the restriction over Ho(fi) °f a functional linear continuous over H^fi). With g and L as in (4.156) and e > 0, we now consider the following particular Fourier problem: Find ut e H:(Q) such that I [(SVu£) • Vu + aouev] dx + e- uev dT Jn Jr = L(v) + - f gv dT, e Jr
VveH\O).
(4.157)
Remark 4.15. If A" is symmetric, there is equivalence between (4.157) and the following minimization problem:
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
359
Find ne eif'ffl) such that J£ut) < J£v),
V v e H\Q),
(4.158)
where Jlv) = \ f [_(X\v) • \v + a0v21 dx - L(v) + i - f (v - gf dT. z Jn ze J r
(4.159)
Thus (4.157), (4.158) is obtained from (4.156) by using a penalty procedure to handle the boundary condition you = g (see Chapter I, Sec. 7, for more details on penalty procedures). The relationship between (4.156) and (4.157) follows from: Proposition 4.14. We suppose that X satisfies (4.47) and that a0 satisfies a0 £ L°°(Q), ao(x) > oc0 > 0 a.e. on Q,
(4.160)
(or possibly—if Q is bounded— a0 e L°°(Q), ao(x) > 0 a.e. on Q).
(4.161)
We then have lira \\ue - u\\HHa)
= 0,
(4.162)
where u (resp. ue) is the solution of the Dirichlet problem (4,156) (resp., of the Fourier problem (4.157)). PROOF. (1) A priori estimates for {u£}E. Let us define a{-, •) bilinear continuous from H\a) x H\Q) into R by
a(v, w)= f l(KVv)-Vw + aovw']dx,
Vi),weff'(Q);
(4.163)
•Jn
taking v = us — u in (4.157), and using the notation of (4.163), we obtain a(ue, us-u)
1 f
+ - \(ue-
gf dT = L(«E - u),
which, in turn, implies
a(uE-u,uc-u)
+ -\\uc-g\\i2ir)
= L(uE-u)-a(u,us-u),
V e > 0. (4.164)
Using (4.52) (which still holds), from (4.164) we obtain a{uE -u,uc-u) where
+ - ||uE - g\\lHT) < C^u, - u\\HHa),
(4.165)
360
App. I A Brief Introduction to Linear Variational Problems
If (4.160) holds, from (4.165) we obtain
Ik "'
ll" " ' " ' - m i n ( a ,
v£
«0)
> °-
<4-166>
If Q. is bounded and if (4.161) holds, we first observe that (4.165) can also be written as a(us - u,ue-
u) + \\uE - u\\lHT) H
\\uc - «||i2 ( r ) < Ct\\ut - u\\BHa);
(4.167)
then, since (from Lemma 4.1 and a > 0) I 1/2
\Vv\2dx+
\\v\\22(r)
defines, over H^Q), a norm equivalent to the usual H1(Q) norm (4.40), we have from (4.167), K - M||Hi(n) < C 2 ,
V E, 0 < £ < 1.
(4.168)
Both (4.166) and (4.168) imply the boundedness of {«J £ in H^il).
D
(2) Weak convergence of {ue}e. From the boundedness of {u8}e in H\Q), we have the existence of u* e H^Q) and of an extracted subsequence—still denoted by {«J£—such that lim uz = u* weakly in H^fi);
(4.169)
£->0
from the continuity of y0 from H1(Q) into L 2 (F) and from (4.169), it then follows that lim y0u8 = you* weakly in L\T).22
(4.170)
Since y o » = 0 , V c e HQ(Q), we have (from (4.157)) a(uc, v) = L(v),
V u e Hj(Q),
V e > 0,
which, combined with (4.169), implies that at the limit, a(u*, v) = Uv),
V v e Hj(Q).
(4.171)
ToshowthatM* = w,itsufficestoshow(from(4.156),(4.171))thaty0«* = g;the boundedness of {uc}e in H 1(fi) and (4.164) imply lim \\u8 - g\\L2{r) = 0,
which combined with (4.170) implies you* = g. Thus we have proved that u* = u (since (4.156) has a unique solution, the whole {u8}8 converges to u). (3) Strong convergence of {u£}8. From the weak convergence of {ue}e to u in H1(Q), we observe that lim {L(u, - u) - a(u, ue - u)} = 0. E->0
22
If F is bounded, we have, in fact, limE_0 ||yo(M,. — u*)\\Lzir) = 0.
(4.172)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
361
Since a(v, v) > 0, V v e H^O), it follows from (4.164), (4.172) that lim a{ut -u,u,-u)
= 0,
(4.173)
lim - ||u£ - g\\ii<x) = 0.
(4.174)
If (4.160) (resp., (4.161)) holds, the strong convergence property (4.162) follows from (4.173) and from a(ue - u,uB- u)> Min(a, a o )||u t -
u\\j,Htl)
(resp., from (4.173), (4.174), and from the fact that
O
a
\\v\2 dx +
r
V12
\v\2 dT) hi
defines a norm equivalent to the //'(Q^norm).
•
Remark 4.16. It is possible to prove Proposition 4.14 using Theorem 7.1 of Chapter I, Sec. 7.3; we should take V = H\Q), a(-, •), and L(-) as above, and j and K defined by
K={v\veH1(Q),yov
= g},
respectively. Remark 4.17. Let us consider the mixed boundary-value problem (4.132) (of Neumann-Dirichlet type), whose variational formulation is given by (4.140). In fact, the solution of this problem can also be obtained (if the usual hypotheses on A, a0 hold) as the limit, as e —> 0, of the solution uE of the following problem of Fourier type: Find ue e H1(Q) such that l(K\u£) • \v + aouv] dx + + - [ govdT,
uev dT = \ fv dx + V » e H\O).
We observe that - f vw dT = [ kvw dT, e Jr 0 Jr
V v, w e H^Q),
where k is defined by k(x) = - i f x e r o , 0 e
if
g^v dT (4.175)
362
App. I A Brief Introduction to Linear Variational Problems
The proof of the convergence of the solution u£ of (4.175) to the solution u of (4.132), (4.140) is left to the reader as an exercise. The above approximation properties are, in fact, of practical interest since many modern finite element codes for solving elliptic partial differential equation problems compute the solution of the (discrete) Dirichlet, Neumann, Neumann-Dirichlet, etc. problems via discrete Fourier formulations (derived from (4.157), (4.175)) with e > 0 "very small." In the finite element code, MODULEF, for example, one uses e = 10~ 30 (see also Perronnet [1] for more details about MODULEF-and further references). From the practical interest of these approximations by penalization of the Dirichlet boundary condition y0 u = g on F (or F o ), a quite natural problem which arises is to estimate the approximation error as a function of s; the following two propositions give partial answers to this problem: Proposition 4.15. We suppose that the hypotheses ona0,A are those of Proposition 4.14 and that L(v) = f fvdx,
V c £ H\n),
(4.176)
•Jn
where fe L 2 (Q). We also suppose that the solution u of (4.156) satisfies (K\u) • n G L 2 (F).
(4.177)
We then have I k - «l| H . (n) = 0{yfe\
(4.178)
where u (resp., u j is the solution of (4.156) (resp., (4.157)). PROOF. Let us define X e L2{Y) by X = (SVu) • n; since u is a solution of (4.156), we have (cf. Sec. 4.3) - V • (KXu) + aou = f in
fl.
(4.179)
1
Multiplying (4.179) by v e H (Q) and using the Green-Ostrogradsky formula (4.33), we obtain a(u, v ) = [ f v d x +
[ XvdT,
V C E H\Cl),
(4.180)
where a(-, •) is still defined by (4.163). We also have a(uE, v) + ~ \(uc-
g)v dV = f fv dx,
V v £ H\Q).
(4.181)
Taking v = uc — u in (4.180), (4.181), by subtraction (also taking into account that y0 u = g), we obtain a(uc - u,ue-
u) + - \ (ue - gf dT = -
l{ut - g) dY.
(4.182)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
363
Since a(v, v) > 0, V I I E H ' P , from (4.182) and from the Schwarz inequality in L2(T) we obtain (4-183)
Ik - dWm-n < eUlvHnCombining (4.182) and (4.183), we obtain, V e, 0 < e < 1, J |ti, - s i 2 rfF < e\\X\\2Hr)
a(ue -u,u.-u)+
which clearly implies (4.178) (if (4.160) holds, we have 1/2
<*o)
Proposition 4.16. The hypotheses on a0, A, L( •) being those of Proposition 4.15, we suppose that the solution u of (4.156) satisfies23 )-
(4.184)
ll«« - «II*'(Q) = 0 ( 4
(4.185)
We then have
where u (resp., u£) is the solution o/(4.156) (resp., (4.157)). PROOF. With I = (5Vu) • n, we define u 1 e H^Q) as the (unique) solution of the Dirichlet problem V • (SVu1) = 0 in n, u 1 = - A on F;
(4.186)
we now define w£ e H1(Q.) by w6 = « + eu1.
(4.187)
We clearly have K
- «IIH.(O) = O(e).
(4-188)
From (4.188) then it follows that to prove (4.185) it suffices to prove that IK - ue\\HHn) = O(e). Taking v = we — u£ in (4.157), we have (with a(-, •) still defined by (4.163)) a(ue, we - uE) + -
(u£ - g)(ws - uc) dT =
f(ws - w£) dx,
23 We have y0H\Q) = {n\n e L 2 (r), 3 p. e H^O) such that n = yop); actually y0H\Sl) = H1I2(T) (for the definition and properties of the Sobolev spaces Hs with s e R, see, e.g., Adams [1], Lions and Magenes [1] and Necas [1]).
364
App. I A Brief Introduction to Linear Variational Problems
which we rewrite 1 r
a(w£ — u£, ws — M£) + (wE — uc)2dT e Jr
= a(w8, w , - « , ) + - f (W, - a Xw. - ut) dT -
f /(w, - ti.) dx. (4.189)
Combining (4.186), (4.187), and (4.189), we obtain 1 r a(ws — us, ws — u£) H— (we — ME)2 d r = e^u 1 , wc — wf.) e Jr
+ L u , we - ««) - J ^(wE - a.) d r - | /(HI, - «J rfxj; (4.190) from (4.180), the second term in the right-hand side of (4.190) vanishes; we then have a(ws — ue, ws — us) + - (ws — us)2 dT = m(ul, we — ue) e Jr which implies a(w8 — uc, wE — u.) + - (w. — u,)2 dT £ Jr < eMax(||X|| l0 o (n) , \\ao\\L^\W\\mm\\wE
- uc\\ma>.
(4.191)
By a now standard reasoning, (4.191) implies that IK - «J H i ( n ) = O(e), which completes the proof of the proposition.
•
Remark 4.18. The hypotheses (4.177), (4.184) concerning the regularity of (AVM) • n on F are often encountered in practice. If the a y , / , g are sufficiently smooth functions of their respective arguments, and if F is a sufficiently smooth curve or surface, properties (4.177), (4.184) will follow from regularity properties of the solutions of the Dirichlet problem (4.156) discussed in, e.g., Necas [1]. 4.5. Finite element approximations of the Neumann, Dirichlet, and Fourier problems
4.5.1. Synopsis We now consider the finite element approximation of the second-order elliptic problems discussed in Sees. 4.2, 4.3, and 4.4; actually the finite element techniques to be described in this section will also be used to approximate the second-order elliptic problem discussed in Sec. 4.6 (and have been used all along in this book to solve nonlinear problems much more complicated than those linear problems discussed in this appendix).
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
365
The following subsections cover only a small part of this very important subject; for more details (theoretical and practical) concerning finite element approximations, see Aubin [2], Ciarlet [l]-[3], Aziz and Babuska [1], Raviart and Thomas [1], Oden and Reddy [1] ,Mercier [1], Strang and Fix [1], Zienkiewicz [1], etc.. 4.5.2. Basic hypotheses: triangulations ofQ and fundamental discrete spaces For simplicity we suppose that Q is a bounded polygonal domain of U2; we then define a family {&~h}h of triangulations of Q such that: (i) (ii) (iii) (iv)
V h, 9~h is a finite collection of closed24 triangles contained in Q; u Terh T = Q; h is the maximal length of the edges of the T e ^ j ; if T, T e .Th, T # T, we have
fnf'
= 0
and either T n V = 0 or T and V have in common a whole edge or only one vertex. Various triangulations for which conditions (i)-(iv) hold are shown in several chapters of this book (see also Appendix II). From 2Th we now define the two following finite-dimensional spaces: Hl = {vh\vheC°(Q),
vh\TePu
Hhh={vh\vheH}:,
VTe^},
wfc = 0 o n r } ,
(4.192) (4.193)
where P1 is the space of the polynomials in two variables of degree less or equal to 1 (i.e., if q e Pu then q{xr, x2) = ax t + fSx2 + y, a, fi, y s U). Let Svh be the gradient of vh (in the sense of distributions); it can be shown that if vh e Hi, then V»*=
Z V(cM7.)Zf
(4.194)
Tef,,
where xt is the characteristic function of the interior T of T i.e., fl
ifxef,
From (4.192), (4.194) it then follows that \vh is piecewise constant over Q, and also that Hhh = Hi n Hl0(Q).
24
i.e., if T €&~h, then T = f u dT (with f = interior of T, dT boundary of T).
366
App. I A Brief Introduction to Linear Variational Problems
Let Efc (resp., HOh) be the (finite) set of the vertices of ^ (resp., of the vertices of 2Th which do not belong to F); we denote by Nh (resp., NOh) the number of elements of 2,h (resp., T,Oh) and we suppose that \,
£» = ZOh u {Qi}^Noh+1
(4.195)
We recall that any polynomial of P1 is entirely defined by the values it takes at the three vertices of a triangle; furthermore, if p, qe Pl and satisfy p(x) = q(x), p(x') = q(x'), where x / x', then p and q coincide on the line xx'. From these properties it follows that T.h (resp., ZOft) is Hj, (resp., Hgh) unisolvent (in the sense of Ciarlet [1], [2], [3]), since V a = {a,}; e UNh (resp., R*0*), there exists a unique vh e Hi (resp, H^) such that vh(Qi) = cth
V i = 1,..., Nh (resp., V i = 1 , . . . , NOh). (4.196)
To each Qt e EA we now associate the function w; (uniquely) defined by w; e Hi,
W; (6 ; )
= 1,
Wi(Qj)
= 0 if i # j .
(4.197)
It is then quite easy to show that &h and ^ 0 / l defined by 0* = {w,}^!,
(4.198)
@Oh = { w j ^ i
(4.199)
are bases of Hi and #oft> respectively; we indeed have Nh
»* = !>*(6<)w«,
V^Gtfi,
(4.200)
V»fceHj h .
(4.201)
i= 1
and also N
vu
We observe that w; vanishes outside the union of those triangles of 2Th having Qi as a common vertex. Remark 4.19. In this appendix we consider piecewise linear approximations on triangles only; for more sophisticated methods making use of higherdegree polynomials, quadrilateral elements, curved elements, and also for finite element methods for three-dimensional problems, see the references given in Sec. 4.5.1. 4.5.3. Some fundamental results on finite element approximations 4.5.3.1. Generalites: synopsis. In this section (Sec. 4.5.3, which closely follows Ciarlet [1], [2], [3] and Mercier [1]), we shall give some results about the
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
367
approximation properties of the particular finite element method introduced in Sec. 4.5.2. We shall be particularly concerned with providing estimates for the finite element approximation errors; we shall use Sobolev norms to express these estimates. 4.5.3.2. Definition and some basic properties of the Sobolev spaces Wm' P(Q). From the importance of Sobolev spaces for estimating finite element approximation errors, we shall complete Sec. 4.2.2 with the following definitions and properties. Let m be a non-negative integer and p satisfies 1 < p < + oo; with Q a domain of UN, we define the Sobolev space Wm'p{Q) as Wm' P(Q) = {v | d*v e Z/(Q),
V a, | a | < m}.
(4.202)
In (4.202), a = {au . . . , a.N} is a multiindex such that V i = 1 , . . . , JV, a; is a non-negative integer, |a| = YA=I a;> a n ^ d*v i s defined by d"v = v if | a | = 0,
h
d"v = —
—— if | a | > 1;
(4.203)
in (4.203), the derivatives are taken in a distribution sense. If m = 0, we clearly have, from (4.202), (4.203), W°-P(D.) = L P (Q); if p = 2, one usually uses the notation Hm(Q.) = Wm-2(Q). It can be proved that Wm- P(Q) is a Banach space if equipped with the norm defined by
(
N
\ ot
Y
p
llr) j>ll
II
0<|a|<m
(A ?(W\
/
For simplicity, one quite often uses the notation INL,p,n= \\v\\w-.p(a), and also /
(4- 205 )i \I/P
Z l|3 a »||£,( n) | =m
•
(4.205) 2
/
If p = + oo, we should use the notation II/J!
^= IIDII
^^
niov ll/5^7jll a 0 < | a | <m
kL.oo.n = max||dai;||Loo(n). \a\=m
A very important property of the spaces Wm' P(Q) is given by the following:
368
App. I A Brief Introduction to Linear Variational Problems
Proposition 4.17. Let Q be a bounded domain of UN; then if F ( = dQ) is sufficiently smooth,25 and if m>~,
(4.206) P
we have Wm'"(Q) c C°(Q),
(4.207)
the injection from Wm'p(Q) into C°(Q) being continuous.26-21 The above proposition is proved in, e.g., Adams [1]. We observe, most particularly, that if N = 2, then WUp(Q) c^C°(Q)
W2p(Q)^C°(U)
if p > 2,
if p > 1.
(4.208)
4.5.3.3. Error estimates using Sobolev norms. (I) Local estimates. Let T be a triangle; we define hT and pT by hT = length of the largest edge of T,
(4.209)
and pT = supremum of the diameters of all the disks contained in T,
(4.210)
respectively (we clearly have h = max T E i f | i hT). Since T is a polygonal domain of U2, from Proposition 4.17 we have Wm'p(t)
ifm>2. P
c, C°(T)
(4.211)
We denote by Ahi = 1,2,3, the vertices of T, and we define nT: C°(T) -> P } by nTv e P , , V U G C°(T), (nTvXAd = v{Ad, V i = 1, 2, 3; (4.212) nT is the linear interpolation operator associated with the three vertices of T. Concerning the approximation properties of nT, we have the following: Proposition 4.18. We consider p,q>
1, and m a non-negative integer such that
2p
W ' (t)
cj C°(T),
2 P
(4.213)
mq
W - (f) c; W ' (f).
(4.214)
Then there exists C independent of T such that llph
l\v\2^^,
V ve W2-P(t).
(4.215)
PT
25
Q can be a polyhedral d o m a i n , for example. In fact, it is compact. In the sequel, we shall use the symbol c? to denote the inclusion, with continuity of the injection. 26
27
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
369
Remark 4.20. In the particular case where q = p, (4.215) holds if m = 0, 1, 2, and then (4.215) reduces to VveW2-"(t)
\v-nTv\m,PtT
ifm = 0, 1,2.
PT
(4.216) 4.5.3.4. Error estimates using Sobolev norms. (II) Global estimates. We use the notation of Sec. 4.5.2; in Sec. 4.5.2 we have introduced a family {&~h}h of triangulations of the domain O. This family {^~h}h is said to be regular if there exists a constant a such that — < a,
VTe£Th,
V h.
(4.217)
PT
Remark 4.21. Condition (4.217) is equivalent to the following property The angles of &~h are bounded from below by 60 > 0, independent of h. (4.218) We now define the interpolation operator nh: C°(fi) -> Hi by nheHl
\/veC°(Q),
(nhv)(Qd = v(Qd,
V0,eZ f c ;
(4.219)
from (4.192), (4.193) we clearly have nhveHhh,
V t i e C°(U),
V = 0 on T.
(4.220)
We observe that nhv\T = nT(v\T),
\fveC\U),
VTe^,
(4.221)
where nT is defined by (4.212); we also observe (from (4.208)) that nhv can be defined if either v e WUp(D.) with p > 2 or v e W2'"(Q) with p > 1. Combining the above properties with the local estimates of Sec. 4.5.3.3 (most particularly (4.216)), we should prove Theorem 4.1, which provides global interpolation error estimates: Theorem 4.1. Suppose that {3~h}h is a regular family of triangulation (i.e., (4.217), (4.218) hold). We then have the following interpolation errors estimates: If p > 2 and m = 0, 1, then \\v-nhv\\mp
VeeW^'Cn),
(4.222)
yvsW2'p(Q),
(4.223)
If p > I and m = 0, 1, then \\v-nhv\\m,p,n
370
App. I A Brief Introduction to Linear Variational Problems
4.1. Suppose that in addition to (4.217), (4.218), the family also satisfies the following property (which in fact implies (4.217), (4.218)):
EXERCISE
k
V/z,
min hT
(4.224)
where fl is a constant. Then prove that there exists a constant C independent of h such that l»*li.p.n^T-|»*lo,p,n. j \
Vcte//i.
(4.225)
[Relation (4.225) plays a very important role in many theoretical questions concerning the finite element approximation of important problems (like the Stokes and Navier-Stokes problems discussed in Chapter VII, Sec. 5).] 4.5.3.5. Some approximation properties of C2 functions. Let T be a triangle of R2 whose vertices are denoted by Au A2, A3 (with At = {aa, ai2}). With nT as in (4.212), we shall estimate, by elementary methods, for v e C2{T) and m = 0, 1.
\KTV - u|m>00>T
For more results in this direction, see e.g., Ciarlet and Wagschal [1]. Let » e C ° ( r ) and consider the linear interpolate nTv; we have (with v, = viAd) 3
nTv(x) = X>jWj(x),
VxeR2,
(4.226)
where the functions w; are (uniquely) defined by Wj e Pu
Wi(Ai)
= 1, w£Aj) = 0
if; / i, i = 1, 2, 3.
(4.227)
Since the functions x -» I , x - * x 1 , x - » x 2 belong to P 1 ; we also have
£w,(x) = l,
Vx,
(4.228)
i= 1
*i= ZatMx),
(4.229)
i= 1
x2=I,at2wl{x).
(4.230)
Properties (4.228)-(4.230) justify the fact that the w;(x) are very often called the barycentric coordinates of x e R2, with respect to Au A2, A3. Another property of the wt is w,(x) > 0,
V i = 1, 2, 3 <* x e T.
(4.231)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
371
Suppose that v e C2(T); we denote by D2v the Hessian matrix of v, i.e.,
fa
d2v
dx2
dx, dx22
8v
8v
J2
-i 8x2
'2
|;
(4.232)
dx2,
we then define \D2v{x)\ by I D2v(x)£. I
\D2v(x)\ = Max !—-A-™ ?€K2-{0}
(4.233)
Isl
(with \%\ = V ^ F + H if ^ = {£i, ^2}); \D2v(x)\ is clearly the spectra/ radius of D2v(x), and since t> e C 2 (T), we have ||£>2*;||0j00,T = Sup ID2v(x)I < +00.
(4.234)
We now prove a first approximation result: Theorem 4.2. Suppose that v e C 2 (T); we then have K x ) - jr r «(x)| < i^||D 2 »|| 0 > 0 0 > r .
(4.235)
xeT
PROOF. Using a second-order Taylor expansion, we have Vi
• (D2v(idxAd
= viAt) = i>(x) + V^(x) • £ 1 ; + 0n
(4.236)
where £, e [x, ^4,] (therefore £; e T if x e T); combining (4.226) and (4.236), we obtain 3 / 3 \ nTv{x) = X f,w;(x) = v(x)[ X w,-(
>
/
3
V=i
i= i
which, if one takes (4.228)-(4.230) into account, implies nTv(x) - v(x) = 1l-Y *>f?teXi • (D2v(Qxld-
(4.237)
t=i
If we suppose that x e T, from (4.228), (4.231), (4.233), (4.234), we have |7tri<x) - v(x)\
W
^^P^IIo.co.rMaxl^l2.
(4.238)
Since Max,- \xA,\ = hT, from (4.238) we obtain -o(x)|^i||fl 2 i7||o,«, r '»f, which finally proves (4.235).
VxeT,
V v e C2(T), •
372
App. I A Brief Introduction to Linear Variational Problems
EXERCISE 4.2. Prove that ll^rf - ^llo.oo.r < hT\v\UaD,T,
V v e C\T)
(4.239)
(we recall that \v\1>a>tT = M a x x e T | \ v ( x ) \ ) . We now consider the more complicated problem, which is obtaining estimates for \\\(nTv — U)|IZ.»(T)XL«(D ( = \nTv ~ ^li.oo.r); m this direction we have the following:
Theorem 4.3. Suppose that v e C2(T); we then have \nTv — v\i „
T
< 3 .
T
-- ||ZD2f||0
T,
(4.240)
sin 0T where 9T is the smallest angle of T. PROOF. From nTv = YI=i viwiiwe
have V(nTv) = X y i V v v i '
which implies—combined with (4.236)—that
VMx))
=
v(^,{ + \ I Vw; (.^4, • (D'KCiM,)).
(4.241)
From (4.228) we have that V(£? = 1 wt) = VI = 0; now using elementary algebra we should prove that 3
->
->
- •
-+
X Vw,(Vti(x) • X/4;) = Vu(x).
(4.242)
From these relations, (4.241) reduces to V(nTv — v)(x) = - YJ Vwj(x^4j • (D2v(^i)xAi)),
(4.243)
which, in turn, implies
\\(nTv - v)(x)\ < UY IVW,| Wax | V x e T.
(4.244)
/ To obtain (4.240), we must estimate YJ= I I Vwf |; actually we have (with mT = measure ofT) -
1
(A ?45t
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
373
From (4.245) it follows that we clearly have Vi = 1,2,3,
|VWJ|<—,
2mT
which, combined with (4.244), implies \\(nTv - v)(x)\ < \ \\D2v\\a,x,T~, 4 mT But since the relation 4mT > hj sin 9T
Vxe T.
(4.246)
always holds, from (4.246) we finally obtain Max \V(nTv - v)(x)\ i.e., relation (4.240).
•
EXERCISE 4.3. Prove (4.242). EXERCISE 4.4. Prove (4.245).
Theorems 4.2 and 4.3 give local approximation results; now suppose that Q is a polygonal bounded domain of R2 and that {&~h}h is a family of triangulations of Q satisfying hypotheses (i)-(iv) of Sec. 4.5.2. With nh the interpolation operator defined by (4.219), we would like to estimate n v h — v\m, oo,n f° r m = Q,l and v a sufficiently smooth function. Using Theorems 4.2 and 4.3, and also the facts that
h = MaxhT, 60 < Min 6T if {2T^h is a regular family of triangulations of Q (i.e., if (4.217), (4.218) hold), l « U . , i i = Max |(»| T )| m>00 , r> nhv\T = nT(v\T),
V v e Cm(U),
V^eC°(Q),
we should easily prove the following: Theorem 4.4. Suppose that {^,}h is a regular family of triangulations of Q and that the linear interpolation operator nh is defined by (4.219). We then have
a,,n, llo,oo,n, \
h
\
U
a
>
t
a
^ sin
2
v\\Oia)ta,
VteC1^
(4.247)
VB6C2(Q),
(4.248)
VveC2(O)
(4.249)
(\\D22vv\\o, oo,n is defined from (4.234) by replacing T by Q).
374 EXERCISE
App. I A Brief Introduction to Linear Variational Problems
4.5. Prove—if Q <= UN—that l»l2,.,n^P2»llo,»,n^N|»|2>oo,n,
VveC2(U).
(4.250)
4.5.4. Application to the finite element solution of the Neumann and Fourier problems In this section we suppose that Q is a bounded polygonal domain of IR2. 4.5.4.1. Synopsis. The Neumann problem for a large class of second-order elliptic partial differential operators has been denned and discussed in Sec. 4.2. In Sec. 4.4 we have shown that the Neumann problem is in fact a particular case of the so-called Fourier problem denned in Sec. 4.4.2. For this reason, in this section (Sec. 4.5.4) we shall consider only the finite element approximation of the Fourier problem (4.141) and then the convergence properties of the approximate solutions. The convergence of the approximate solutions follows from the general approximation results of Sec. 3, combined with the approximation properties stated in Sec. 4.5.3. 4.5.4.2. Formulations of the continuous and approximate Fourier problems. We recall (from Sec. 4.4) that the operator formulation of the Fourier problem is - V • (AM) + aou = f in Q, (AVH) • n + ku = g on F, (4.251) the corresponding variational formulation being: Find u e H *(Q) such that (AVw) -\vdx + aouv dx + kuv dT n JQ Jr = [fvdx+ \gvdT, Jn Jr
Vcefl'CQ).
(4.252)
We suppose that the hypotheses on A, a0, k, f g made in Sec. 4.4.3 still hold; then problem (4.251), (4.252) has a unique solution. To approximate (4.251), (4.252), we consider a family {&~h}h of triangulations of Q (as in Sec. 4.5.2) and the discrete space Hi defined by (4.192). Then we approximate (4.251), (4.252) by the following finite-dimensional linear variational problem: Find uh e Hj, such that
Jn
(A\uh) • Svh dx +
= {fvhdx+ Jn
aouhvhdx+ Jn
[gvhdT, Jr
problem (4.253) clearly has a unique solution.
Jr
kuhvh dT
\fvheHl;
(4.253)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
375
4.5.4.3. Convergence of the approximate solutions. Concerning the convergence of {uh}h to the solution of (4.252), we have: Proposition 4.19. Suppose that A, ao,k,f, g satisfy (4.146)-(4.149); also suppose that {3Th}h is a regular family of triangulations of Q (in the sense of Sec. 4.5.3.4). We then have lim | K -
UWHHCD
= 0,
(4.254)
where uh (resp., u) is the solution of (4.253) (resp., (4.251), (4.252)). PROOF. TO apply Theorem 3.2 of Sec. 3.3 (with Vh = Hi), it suffices to find TT and rh obeying (3.1). Since (cf. Sec. 4.2.2, Proposition 4.2)
we can take "V = &(Q). We then define rh by rhveHlh,
VveC°(U),
(rhv)(Qd = v(Qd,
^ QteZh,
(4.255)
where (cf. Sec. 4.5.2) Y.h is the set of the vertices of JJ; actually rh coincides with the interpolation operator nh denned by (4.219). Since {^}/, is a regular family of triangulations of Q, it follows from Theorem 4.4 of Sec. 4.5.3.5 (we can also use Theorem 4.1 of Sec. 4.5.3.4) that l i m \\rhv - v\\HHa)
= 0,
V c e
ft->0
which completes the proof of the proposition.
•
The convergence result (4.254) still holds if a0 e LX(Q.),
a0 > 0,
| ao(x) dx > 0
(resp., if a0 = 0, k e L°°(r),
k > 0,
f k(x) dY > 0), Jr
the other hypotheses on A, k, f, g (resp., A, / , g) remaining the same. 4.5.4.4. Formulation of the approximate Fourier problem (4.253) as a linear system. Comments on numerical integration. We retain the notation of Sec. 4.5.2. Let 0&h be the basis of H\ defined by (4.197), (4.198); we then have
376
App. I A Brief Introduction to Linear Variational Problems
where {2j}^=i = £*• Setting iij = uh(Qj), the approximate Fourier problem (4.253) is then equivalent to the following linear system:
f
f (SVw,-) • \wt dx + f aowjWi dx + f /cWj-w, dF L = I fwt dx +
gwt dT,
1 < i < Nh,
(4.256)
whose unknowns are Uj,j = 1 , . . . , Nh. From the properties of A, a0, k, the matrix of the above system is positive definite; it is symmetric if A(x) is symmetric a.e. on Q. From these properties, system (4.256) has a unique solution. The solution of linear systems like (4.256) will be considered in Sec. 4.5.7. What we would like to do in this section is to discuss the computational work required to obtain the above system (namely the associated matrix and the right-hand-side vector). For 1 < (', j < Nh, define atj and /?; by
« y = f[(AVw,;
'; + aowjwi~] dx +
kwjWi dT,
(4.257)
•la.
t=
f fw,dx+ (
Jn
(4.258)
^r
respectively. Then the linear system (4.256) can be written as follows: AfcU, = pA,
(4.259)
where A, = ( a ; i ) U i , j s J V h , U, = {wjf^, pft = {pt}^v To obtain A, and p,, we clearly have to compute a large number of integrals. Define Q{ and Tt (see Fig. 4.3) by Q ; = support (w;),
r ; = 5Q ; ;
the sets Q ; and F,- are small if Nh is large (i.e., if h is small).
Figure 4.3. The sets Q, and rt. (a) Qt is not a boundary node, (b) Qt is a boundary node.
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
377
An important simplification follows from the fact that in (4.257) (resp., (4.258)) we have
L-L.-
Wrrr
(resp.
f = I". \ = f \ hence the above integrals have to be evaluated on a small number (which is zero most of the time) of triangles and edges of !Th. The resulting matrix Ah is a sparse one. Since, in practice, the exact calculation of a y and /?,- is generally impossible, numerical integration has to be used to approximate the above quantities. For simplicity we suppose that A and / (resp., k, g) are continuous over Q (resp., F) (the generalization to piecewise continuous functions is straightforward if we suppose that their discontinuities are supported by edges of 2Th (resp., vertices of 9~h belonging to F)). Let T be a triangle of 2Th contained in Ut n Uj (resp., Q,); the calculation of 7,ij and ^ requires the evaluation of Wi + flowiwi]
dx
(4.260) j
and 'wtdx,
(4.260)2
respectively. Denote the three vertices of T and its centroid by Q1T, Q2T, and Q3T, and GT, respectively. Remembering that \wh \wj are constant over T and concentrating on simple numerical integration formulas, we may approximate the integral in (4.260)! by either meas (T)X(GT)W(Wj\T)
• V(w,| r ) + * meas(T) £ ->
ao(QkT)W]{QkT)Wi{QkT),
k=l
(4-261)! or \ meas(T) ^ {X(QkT)\(wj\T) J
• V(w,| r ) + ao(QkT)wj(QkT)wJiQkT)};
(4.261)2
/t=i
formula (4.261)2 is obtained through the application of the two-dimensional trapezoidal rule to (4.260)^ Similarly, we should approximate the integral in (4.260)2 by \ meas(T) £ /(<2fcrMQ t r );
(4.262)
378
App. I A Brief Introduction to Linear Variational Problems Qt
Figure 4.4
From (4.262) and from the properties of the w; it then follows that j n fw( dx will be approximated by
E where a>, = meas(Q;) and {<2Jf=i = Zfc. We now consider the calculation of the boundary integrals occurring in (4.257), (4.258); we have to evaluate integrals like (see Fig. 4.4)
f
kwt Wj dT,
jQiQj
[
kwf dT,
JQiQj
f
gwt dT.
(4.263)
JQiQj
Again using the trapezoidal rule and remembering that w;(Q;) = 1, w^Qj) = 0 if i / j , for the above integrals we obtain the following approximations:
o, $\&Qj\KQd,
$\&QMQd,
respectively. Using Simpson's rule, more elaborate approximations can be obtained for the integrals in (4.263). EXERCISE 4.6. Prove that the discrete Fourier problem, obtained using numerical integration, is still well posed. Also prove the convergence of the approximate solutions. For more details on numerical integration and its effect on the convergence of finite element methods, see Strang and Fix [1], Ciarlet [ l ] - [ 3 ] , Oden and Reddy [1], and Bernadou and Boisserie [1].
Remark 4.22. Relations (2.261) require a knowledge of the first-order derivatives of the w;. If T is a triangle whose (positively oriented) vertices are denoted by A, B, C, with coordinates {au a2}, {bu b2}, {cu c2}, respectively, and if v is a polynomial of Pi such that v(A) = vA, v(B) = vB, v(C) = vc,
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
379
we indeed have dv dxt
1 {vA(b2 - c2) + vB(c2 - a2) + vda2 - b2)}, 2m(T) (4.264)
dx2
2m(T)
{vA(b1 - cj) + vB(c1 - ax) + vc(ax - bj)},
where
m(T) = meas(T) = 0.5\CA x CB\ = 0.51(0! EXERCISE
Cl){b2
- c2) - (&! -
Cl)(a2
- c2)\. (4.265)
4.7. Prove (4.264), (4.265).
4.5.5. Application to the finite element solution of the Dirichlet and mixed Neumann-Dirichlet problems With Q as in Sec. 4.5.4, we now consider the finite element solution of Dirichlet and mixed Neumann-Dirichlet problems. Since (see Sec. 4.3.4 for more details) the Dirichlet problem appears as a particular case of the other problem, we shall consider the second one only (actually the extension to the mixed FourierDirichlet problem is straightforward). 4.5.5.1. Formulation of the continuous and discrete Neumann-Dirichlet problems. We suppose that F ( = dQ) is the union of F o , Fj such that F o u Fj = F, F o n Fi = 0 ; the continuous problem is then defined by — V • (AVM) + aou = / i n Q,
u = g0 on F o ,
(K\u) • n = g^ on F. (4.266)
With Vo denned by Vo = {v | v e H^fi),
v = 0 a.e. on F o },
for (4.266) we have the following variational formulation: Find u e H^Q) such that u = g0 on F o and I (K\u)-Vvdx •In
+ \aouvdx= Jn
| fv dx + \ gxvdT, Jn
VveV0.
JTi
(4.267) We suppose that the hypotheses on X, a0, f, g0, gt are those made in Sec. 4.3.4; then problem (4.267) has a unique solution. To approximate (4.266), (4.267), we consider a family {&~h}h of triangulations of Q—as in Sec. 4.5.4.2— and we suppose that the following property holds: the points at the interface of F o and Tx are vertices of STh. (4.268) From the above property, the following subspace of Hi, defined by Hl
vh = 0 o n F 0 } ,
(4.269)
380
App. I A Brief Introduction to Linear Variational Problems
is also a subspace of Vo. Eventually we suppose that g0 is continuous over F o (F o : closure of F o in F). As the approximate problem we consider: Find uh 6 Hi, uh(Q) = go(Q), V Q e To n Z_, such that
Jn
(A\uh)-\vhdx
+ Jn
aouhvhdx=
Jn
fvhdx+
g^dT,
\/vheVOh.
JTI
(4.270)
If the above hypotheses hold, the approximate problem (4.270) has a unique solution. 4.5.5.2. Convergence of the approximate solutions. Concerning the convergence of {uh}h to the solution of (4.270), we have: Proposition 4.20. Suppose that A, a0, f, gx satisfy the hypotheses made in Sec. 4.3.4 and that {2r^h is a regular family of triangulations of Q. (in the sense of Sec. 4.5.3.4) obeying (4.268). Also suppose (for simplicity) that g0 = <70|r0> where g0 is Lipschitz continuous over Q. We then have lim \\uh - u\\HHn) = 0,
(4.271)
h->0
where uh (resp., u) is the solution of (4.270) (resp., (4.266), (4.267)). PROOF. We sketch the proof in the case where 0O = O o n r 0 ,
(4.272)
implying that one can take g0 = 0 over fi. If (4.272) holds, problems (4.267), (4.270) reduce to: Find u e Vo, such that f (SVu) • \v dx + [aouvdx= Jn
•'Q
f fv dx + [ g^vdY, Jn
VDGF0;
(4.273)
VvheVoh,
(4.274)
JTi
Find uh e Voh, such that \(Ksiuh)-Svhdx+ Jn
\aouhvhdx= Jn
\ fvhdx+ Jn
\ givhdT, •'ri
respectively. To apply Theorem 3.2 of Sec. 3.3 (with Vh replaced by Voh), it suffices to find y and rh which obey (3.1). Define V by y = {v | v e C'iQ),
v = 0 in the neighborhood of Fo}.
Since F is polygonal and (from (4.268)) F o and Vt are "nice" subsets of F, we have (from, e.g., Necas [1]) F H 1 ( n ) = Vo.
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
381
If v € y and if rh is still defined by (4.255), we have—since (4.268) holds—rhv e Voh; on the other hand, we still have (see the proof of Proposition 4.19 for more details) lim \\rhv - v\\HHa) = 0,
Vve"T,
ft-»0
which completes the proof of the proposition.
•
4.8. Prove Proposition 4.20 if g0 ^ 0, with g0 = go\ro, g0 being Lipschitz continuous over Q.
EXERCISE
4.5.5.3. Formulation of (4.270) as a linear system. We retain the notation of Sec. 4.5.2.28 Let @h be the basis of Hi defined by (4.197), (4.198); we then have
«*= f W e > J - +
t
go(Qj)wj,
(4.275)
where the vertices of yH belonging to F o have been ordered from NOh + 1 to Nh. Therefore it is clear that @Oh = {wj}^\
is a basis of VOh.
(4.276)
Again setting uh(Qj) = Uj, it follows from (4.275), (4.276) that the approximate problem (4.270) is equivalent to the linear system Y, AVWj • Vw; dx + a0WjWi dx \uj j=i\_Jn Jn J =
fw,dx Jn x go(Qj),
+
giwtdr JTI l
-
YJ \\wj-\Widx j=N0h +i\_Jn
+ Jn
a0WjWidx\ J (4.277)
From the properties of A, a0, the matrix in (4.277) is positive definite (it is symmetric if A(x) is symmetric a.e. on Q); from these properties, (4.277) therefore has a unique solution. All the comments and observations, made in Sec. 4.5.4.4 about the necessity of using numerical integration to obtain a computationable approximation of (4.277) still hold. 4.5.6. Some comments on the relationships between finite element and finite difference methods It is well known (by the specialists, at least) that most finite difference methods can be viewed as particular case of finite element methods (and vice versa). To justify this claim, we consider the following Dirichlet problem: - AM = / in Q,
u = g on T,
(4.278)
28 Except for Noh, which is now defined by Noh = dim Voh = number of vertices of ^ which do not belong to r 0 .
App. I A Brief Introduction to Linear Variational Problems
382
Figure 4.5
where Q = ]0, l [ x ] 0 , l[, r = dQ, where / e C ° ( Q ) , and g e C°(T) with g = g\T, where g e H^Q). If the above hypotheses on / and g hold, then the Dirichlet problem (4.278) has a unique solution in H^Q). To approximate (4.278), we consider a positive integer JV and define h by h = 1/(JV + 1). The triangulation 3~h used to approximate (4.278) is as in Fig. 4.5 (where N = 3). The vertices of 5~h are the points Qtj defined by Qij = {ih,jh},
0
< N + 1.
Applying the results of Sec. 4.5.5, we obtain (with the obvious notation) the following approximate problem:
O
n
VWy • \wkl dx )ukl =
/
Jn
fWij,
g(Qki),
1 < U j < N,
(4.279)
Qkler\Jn
where wkl is the basis function associated with Qkl by (4.197), (4.198). Using the trapezoidal rule to evaluate \nfwijdx (see Sec. 4.5.4.4), we find (after simplification by h2) that (4.279) reduces to (with ftj = f(Qu)) iJ+l
for 1 < i,j < N,
JiJ
'
with ukl = g(Qkl)
if Qu e T.
(4.280)
Hence, the finite element approximate Dirichlet problem (4.279) reduces to the usual five-point finite difference scheme to approximate the Dirichlet problem (4.278). A similar conclusion holds for more complicated problems than (4.278) (a numerical integration procedure is required in most cases).
4 Application to the Solution of Elliptic Problems for Partial Differential Operators EXERCISE
383
4.9. Prove that (4.279) reduces to (4.280).
4.5.7. Solution of the linear systems equivalent to the approximate problems 4.5.7.1. Synopsis. In the previous sections we have discussed the finite element approximation of a fairly broad family of elliptic boundary-value problems. It has been shown that the discrete problem is in fact equivalent to a linear system whose matrix is positive definite (possibly symmetric) and sparse. Actually the solution of the sparse linear systems, resulting from the finite element discretization of elliptic boundary-value problems by finite element methods, has inspired a large number of papers, and it is impossible to mention all of them. Some references have been given in Chapter VII, Sec. 5.8.7.2, and many other significant references can be found in Duff [2] and George [1] (see also Golub and Meurant [1], Mercier [1], Mercier and Pironneau [1], and Thomasset [1]). In Sees. 4.5.7.2,4.5.7.3, and 4.5.7.4 we shall consider the solution of the linear system Ax = b
(4.281)
by Cholesky, over-relaxation, and conjugate gradient methods, respectively; we suppose that A is a symmetric and positive-definite real matrix. 4.5.7.2. Solution of the linear system (4.281) by the method of Cholesky. We present the standard Cholesky method only (for further comments, see the above references and also Ciarlet [4]). Since A = (fl,-j)i<;,j<jv is symmetric and positive definite, there exists a unique lower triangular matrix L = (hj)i
V i = 1 , . . . , N,
A = LI/.
(4.282)
The elements ltj of the matrix L are given by ltJ = 0 'n=V«ii.
if j > /,
^1=^
(4.283)!
Hi=l,...,N,
(4.283)2
M1
if j > 2, then 7-1
hi = J % " E $> hi = V
*=i
fly ~ E likhk k=l
~
if '• >l
hi
(4.283)3 Once L has been obtained, solving (4.281) is trivial since it is equivalent to Ly = b,
L'x = y,
(4.284)
384
App. I A Brief Introduction to Linear Variational Problems
and both systems in (4.284) are easily solved since L and V are triangular matrices. An important property of the above Cholesky factorization is that it preserves the band structure of A; more precisely, if A is a 2p + 1-diagonal matrix (i.e., atj = 0 if | i — j \ > p), then a similar property holds for L, since— from formulas (4.283)—we have lu = 0 if i — j > p. However, if A is a sparse matrix, L is usually less sparse than A. EXERCISE
4.10. Prove the existence of L obeying (4.283).
4.5.7.3. Solution of the linear system (4.281) by over-relaxation. Since Chapter V is concerned with relaxation methods, in this section we shall only give some brief comments on their applications to the solution of linear systems like (4.281). With co a positive parameter, we consider the following iterative method
(where b =
iU
? x° = {x°u ...,x°N}
given arbitrarily;
(4.285)
then for n > 0, x" being known, we compute x" + 1 component by component, by
1
(
W aa
fori=l,...,
N.
If co > 1 (resp., co = 1, co < 1), algorithm (4.285), (4.286) is an over-relaxation (resp., relaxation, under-relaxation) method. From Chapter V, Sec. 5.4 it follows that algorithm (4.285), (4.286) converges, V x ° e R J V , t o x = A^bifandonlyifO < co< 2. For an analysis of the speed of convergence see, e.g., Varga [1], Young [1], Isaacson and Keller [1], and Ciarlet [4]. The choice of a> in algorithm (4.285), (4.286) may be critical, and in general the optimal value of this parameter is unknown. However, there exists a method, due to D. Young, which allows the automatic adjustment of this parameter; theoretically this method, discussed in Varga [1] and Young [1], applies to a class of matrices less general than symmetric positive-definite matrices; however, we observed that it behaved quite well on some problems for which the hypotheses stated in the above references were not fullfilled. 4.5.7.4. Solution of the linear system (4.281) by a conjugate gradient method. The solution of linear and nonlinear problems by conjugate gradient methods has been discussed in several places in this book, and several references concerning these methods have been given. Concentrating on the solution of the
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
385
linear system (4.281), a standard algorithm is as follows (with S a symmetric positive-definite matrix): x°° e UNN arbitrarily given,
(4.287)
r° = Ax° - b,
(4.288) (4.289) (4.290)
w° = g°.
Then, for n > 0, assuming that x", r", g", w" are known and that r" # 0, we obtain xn+1,r"+1,gn+1,wn+1 by
£&
(4291)
(4.292) r" +1 = r " If r"
+1
= 0, then x"
+1
pnAwn.
(4.293)
+
= x = A" *b; if r" * # 0, then compute g»
+1
=S-1r"+1, +1
7
+1
(r" ,g" )
"
+i=
^TI^'
w n+i = g " + i
+ y n+lW ".
(4.294)
(4295) (4.296)
Then n = n + 1 arcd go to (4.291). In (4.287)-(4.296), (•, •) denotes the usual scalar product of UN (i.e., (x, y) = X f = 1 x i y i i f x = {x i }f = 1 ,y={^}f = 1 ). If roundoff errors are neglected, then, V x 0 , there exists n0 < N such that x"° = x = A" 1 b. The above matrix S is a scaling (or preconditioning) matrix; a proper choice of S can accelerate the convergence in a very substantial way (clearly, S has to be easier to handle than A). It is clear that if one uses a Cholesky factorization of S, it will be done once and for all, before running algorithm (4.287)-(5.296). The solution of linear systems of equations, with A nonsymmetric, by methods of conjugate gradient type is discussed in, e.g., Concus and Golub [1] (see also Duff [2] and George [1]). 4.6. Application to an elliptic boundary-value problem arising from geophysics 4.6.1. Synopsis: formulation of the problem In this section we shall discuss the solution, by the methods of the previous sections, of a linear elliptic boundary-value problem originating from geophysics. More precisely, this problem occurs in the modelling of the interactions between jets of electrons and protons emitted by the Sun and the Earth
386
App. I A Brief Introduction to Linear Variational Problems
magnetosphere (see Blanc and Richmond [1] for more details). After an appropriate change or coordinates, the above problem takes the following formulation (where u is an electrical potential and Q = ]0, 1 [ x ]0, 1 [): Find a junction u satisfying - V-(AVw) = /m i/O <
= go(xi)
I)
(K\u • n)
Xi
1, x2) -= 0 ,8u
* U (
x
dx1 \ A i)
3
v-
(4.297)2
<1,
(4.297)3
i/O <; x 2
M(0, X 2 ) = «(1, x 2 )
n)(0, x 2 ) +
(4.297)x
«i,0)j = (,
I/O
< x 2 < 1,
(4.297)4
i/O < x,
OXi
(4.297)5 (4.297)6
w(0, 0) = i«(1, 0);
in the above relations, A, / and go,g1,k are given functions, denned over Q and F, respectively. Actually, if A = (a^! <;_ j i 2 , we have, for (4.297)4 and (4.297)5, the following more explicit formulations: du
du \
I ',x2)=
du
du
ati^r + ai2^-
(4.298) )(xlt 0) - —- k{xx) — - x 1( 0) 5x \ 5x = gr1(x1)
ifO <x1 < 1. (4.299)
The boundary conditions in problem (4.297) may seem rather complicated; actually, using an appropriate variational formulation, it will be seen in the following sections that problem (4.297) is almost as easy to solve as the Dirichlet, Neumann, and Fourier problems discussed earlier. 4.6.2. Variational formulation of problem (4.297) Let v be a smooth function denned over Q (we may suppose that v e C°°(Q), for example). Multiplying (4.297)t by v and applying the Green-Ostrogradsky formula (4.33), we obtain
f (K\u) -\vdx= Ja
f fv dx + f (K\u) • nv dr. Jn
(4.300)
Jr
Now suppose that v(xu 1) = 0 i/O < x t < 1,
v(0, x2) = v(l, x2) i/O < x 2 < 1; (4.301)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
387
it follows from (4.301) and (4.297)4 that the boundary integral in (4.300) reduces to
f
(K\u)-nvdr,
where F, = {x|x = {x 1 ; x 2 }, 0 < x, < 1, x 2 = 0}, implying, in turn, that (4.300) reduces to
f (K\u) -\vdx=
f fv dx + f (K\u) • nv dT.
(4.302)
Jn Jn Jrt Combining (4.297)5 and (4.302), and using the second relation (4.301), we obtain (after integrating by parts over I \ ) f (XVu) -\vdx+ Jn = [ fvdx+
f k(x,) ~ (x,, 0) -^- (x,, 0) dx, Jo ox, ox, [ g,v dT.
(4.303)
Conversely, it can be proved that if (4.303) holds for every vsV,
where
X
r = {v\v £ C (Q), v(0, x 2 ) = v(l, x2) if 0 < x 2 < 1, v = 0 in the neighborhood of F o }
M. 304^
where T o = {x|x e {x1; x 2 }, 0 < x, < 1, x 2 = 1}, then u is a solution of the boundary-value problem (4.297). Relations (4.301), (4.303) suggest the introduction of the following subspaces of H 1 (Q): V = {v\v e H^Q), v(0, x2) = v(l, x 2 ) a.e. 0 < x 2 < 1, (d/dx,)v(x,, 0) £ L 2 (0, 1)}, Vo = {v\v e F, t;(x1; 1) = 0 a.e. 0 < x, < 1}.
(4.305) (4.306)
Suppose that V is endowed with the scalar product (v, w)v = (v, w)HHn) +
— v(x,, 0) — - w(x1; 0) dx,
Jo Q-X,
ax,
(4.307)
and the corresponding norm \\v\\r = (v,v)r12.
(4.308)
We then have the following: Proposition 4.21. The spaces V and Vo are Hilbert spaces for the scalar product and norm defined by (4.307) and (4.308), respectively. Moreover, the seminorm 1/2 l\\Vv\2dx+ C dv u 0) dx, \Jn
Jo
defines a norm equivalent to the V-norm (4.308) over Vo.
388
App. I A Brief Introduction to Linear Variational Problems
4.11. Prove Proposition 4.21. We now define a bilinear form a: V x V -* U and a linear functional L:F-+Rby
EXERCISE
a(v, w) = f (K\v) -\wdx+ Jn
f k(Xl) • / - !<*!, 0) - ^ - W(JC1; 0) dx ls Jo dxt dx t (4.309)
L(t>) = f / y r f x + f s ^ d T , (4.310) Jn Jr t respectively. We suppose that the following hypotheses concerning A, k, f, 0i hold: /eL2(Q),
^ E L
2
^ ) ,
(4.311)
k e L°°(0, 1), k(xx) > a0 > 0 a.e. on ]0, 1[,
(4.312)
X satisfies (4.47).
(4.313)
From the above hypotheses, we find that a(-, •) is bilinear continuous over V x V and Vo-elliptic, and that L(-) is linear continuous over V; we can therefore apply Theorem 2.1 of Sec. 2.3 to prove: Proposition 4.22. If the above hypotheses on A, k,f, g± hold, and if g0 in (4.297)j satisfies do = 0o | r0
with
So e V,
(4.314)
\/veV0
(4.315)
then the linear variational problem: Find usV such that u \To = g0 and a(u,v) = L(v),
has a unique solution; this solution is also the unique solution in V of the boundaryvalue problem (4.297). PROOF. Define u e Vo by u = u — g0; u—if it exists—is clearly a solution of the following linear variational problem in Vo:
ueV0, a(u,v) = L{v)-a(go,v).
(4.316)
Since v —> L(v) — a(g0, v) is linear and continuous over Vo, and since a{-, •) is F0-elliptic, it follows from Theorem 2.1 of Sec. 2.3 that (4.316) has a (unique) solution, in turn implying the existence of u solving problem (4.315). The above u is clearly unique, since if ux and u2 are two solutions of (4.315), then u2 — ux e Vo and also a(u2 — ux, u2 — ut) = 0; the Po-ellipticity of a{-, •) then implies Ui = u2. •
Remark 4.23. Suppose that A is symmetric; this in turn implies the symmetry of a(-, •)• From Proposition 2.1 of Sec. 2.4 it then follows that (4.315) is equivalent to the minimization problem:
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
389
Find ue V,u = g0 onT0, such that J{u) < J(v),
VceF,
v = g0
on F o ,
where
J(v) = -z f (X\v) -\vdx + - f k(Xl) Ja
^ Jo
- f fv dx - f g^ dr. 4.6.3. Finite element approximation of problem (4.297) In this section we consider the approximation of the boundary-value problem (4.297) via the variational formulation (4.315). Actually the finite element approximation discussed in the sequel is closely related to the approximations of the Neumann, Dirichlet, and Fourier problems discussed in Sees. 4.5.4 and 4.5.5. To approximate (4.297), (4.315), we consider a family {3Th}h of triangulations of Q satisfying the hypotheses (i)-(iv) of Sec. 4.5.2 and also: (v) If Q = {0, x2} is a vertex of ,Th, then Q' = {1, x2} is also a vertex of jTh, and conversely (i.e., STh preserves the periodicity of the functions of the space V (cf. (4.305)). With Hj, still defined by (4.197), we approximate the above spaces V and Vo (cf. (4.305), (4.306)) by Vh=VnHl
VOh =Von
= {vh\vh e H{, vh{0, x2) = vh(l, x2) if 0 < x2 < 1}, (4.317) Hi = {vh\vh e Vh, vk(Xl, 1) = 0 if 0 < xx < 1}.
(4.318)
We suppose that g0 (in (4.297)2, (4.314)) also satisfies g0 e C°[0, 1] and (withF 0 = {x\x = {xL, l},0 < Xi < 1}) we approximate the problem(4.297), (4.315) by: Find uh e V such that uh(Q) = go(Q), V Q vertex of^h located on F o and a(uh, vh) = L(vh),
Mvhe VOh,
(4.319)
where a(•, •) and L are still denned by (4.209), (4.310), respectively. We should easily prove that the approximate problem (4.319) has a unique solution if (4.311)-(4.313) hold. 4.12. Prove that the approximate problem (4.319) has a unique solution if the above hypotheses hold. EXERCISE
The convergence of the approximate solutions follows from:
390
App. I A Brief Introduction to Linear Variational Problems
Proposition 4.23. Suppose that the above hypotheses on A, k, f g0, g1 and {^~h)h hold. Also suppose that g0 =
(4.320)
where uh (resp., u) is the solution of (4.319) (resp., (4.297), (4.315)). PROOF. We sketch the proof in the case where g0 = 0 on F o , implying that we can take g0 = 0 over U. Problems (4.315), (4.319) reduce to: Find ueV0 such that a(u, v) = L(v),
V»eF0.
(4.321)
a(uh,vh) = L(vh),
VvheV0h,
(4.322)
Find uh e Voh such that respectively. To apply Theorem 3.2 of Sec. 3.3 (with Vh replaced by Voh), it suffices to find f" and rh obeying (3.1). Define V by (4.304); it has been proved by H. Beresticky and F. Mignot (personal communications) that
If v e y and if rh is still defined by (4.255), we have—since condition (v) on {&~h}h holds— rhv e Voh; on the other hand, we still have (see the proof of Proposition 4.19 for more details) lim \\rhv - v\\HHn) = 0,
Vȣf.
(4.323)
Since 1/2 W v =
- Hxi, 0)
W
L 2 ( O . 1),
it follows from (4.323) that to complete the proof of the present proposition, it suffices to prove that lim
(rhv-v)(Xl,0) d
= 0,
V»ef"
(4.324)
L 2 (0,l)
Let us denote by y^ the function xY -> v(xu 0); we clearly have Vlfrd" -
v
) = sh7yV - JiV,
where sh is the interpolation operator defined as follows: sh
Vtf>eC°[0, 1],
SiXfi) =
for any pair {£u 0}, {^i, 0} of consecutive vertices of STh belonging to
(we recall that Tx = {x\x = {xu 0}, 0 < x, < 1}).
4 Application to the Solution of Elliptic Problems for Partial Differential Operators Since vsY that
391
implies that jiV e C°°[0,1], it follows from standard approximation results
lim fz->0
= 0,
dx1
mo, i
i.e., (4.324) holds. EXERCISE
•
4.13. Prove Proposition 4.23 if g>0 # 0.
Remark 4.24. The approximation by variational methods of problems closely related to (4.297) is considered in Aubin [2]. 4.6.4. Some comments on the practical solution of the approximate problem (4.319) To solve the approximate problem (4.319), we can use a penalty method similar to the one used in Sec. 4.4.4 to approximate the Dirichlet problem by a Fourier one; the main advantage of this formulation is that the discrete space under consideration is still Hi denned by (4.197). A possible penalty approximation of (4.319) is (with e > 0): Find u\ e Hj, such that, V vh e Hi, 1 f1 a(ul vh) + -e («J(1, x2) - ifh(0, x2)Xvh(l, x2) - vh(0, x2)) dx2 Jo If1 If1 + - ul(xu \)vh(xu l)dxj = L(vh) + ~ 0OJI(XIK(XI> !) eJo
£ Jo
dx
u
(4.325)
where gOh e C°[0, 1] coincides with g0 at those vertices of 3~h located on F o ( = {x\x = {xu 1}, 0 < xl < 1}) and is linear (i.e., belongs to P t ) between two consecutive vertices of 2Th located on F o . Such an approximation is justified by lim u\ = uh,
(4.326)
E^O
where uf, and uh are the solutions of (4.325) and (4.319), respectively. 4.14. Prove (4.326). Another possibility is to work directly with the spaces Vh and VOh, taking into account the periodicity conditions uh(0, x2) = uh{\, x2), vh(0, x2) = vh(l, x2) and also the fact that uh(xu 1) = gOh(xi), vh(xu 1) = 0, V vh e VOh. This second approach will require an explicit knowledge of vector bases for EXERCISE
392
App. I A Brief Introduction to Linear Variational Problems
Vh and VOh; obtaining such bases from the basis of Hi defined by (4.198) (see Sec. 4.5.2) is not very difficult and is left to the reader as an exercise. We should again use numerical integration to compute the matrices and right-hand sides of the linear systems equivalent to the approximate problems (4.319) and (4.325). We shall conclude by mentioning that the methods discussed in Sec. 4.5.7 still apply to the solution of the above linear systems.
4.7. On some problems of the mechanics of continuous media and their variational formulations 4.7.1. Synopsis In this section, which very closely follows Mercier [1, Chapter 2] and Ciarlet [2], we briefly discuss some important problems of the mechanics of continuous media which are the three-dimensional linear elasticity equations (Sec. 4.7.2), the plate problem (Sec. 4.7.3), and Stokes problem (Sec. 4.7.4). After describing the partial differential equations modelling the physical phenomena, we discuss the variational formulations and various questions concerning the existence and uniqueness of the solutions of these problems. 4.7.2. Three-dimensional linear elasticity Let Q <= |R3 be a bounded domain. Let T be the boundary of Q and suppose that r = T o u I \ with r 0 , r \ such that jYonrt dF = 0 (a typical situation is shown in Fig. 4.6). We suppose that Q is occupied by an elastic continuous medium and that the resulting elastic body is fixed along F o . Let f = {/J}f=i be a density of body forces acting in Q and g = {gt}f=x be a density of surface forces acting on F1. We denote byu(x) = {Wj(x)}f=i the displacement of the body at x.
Figure 4.6
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
393
In linear elasticity the stress-strain relation is ffy(u) = A(V • u)<5u + 2 ^ » ,
with £ i /u) = \{~^
+^ \
(4-327)
where atj and etj denote the components of the stress and strain tensors, respectively; X and JX are positive constants and are known as the Lamme coefficients. The problem is to find the tensor a = (o-y), the displacement u = {u;}f= x if Q, f e (L 2 (Q)) 3 and g e (L 2 (r\)) 3 are given. The equilibrium equations are JLatj
+ ft = 0
infi, onTh
ut = 0 o n T 0 ,
. = 1,2,3, i = 1,2,3,
i = 1,2,3.
(4.328^ (4.328)2 (4.328)3
We have used the summation convention of repeated indices in the above equations. To obtain a variational formulation of the linear elasticity problem (4.327), (4.328), we define a space V, a bilinear form a(-, •) and a linear functional L by V = {v | v e (if H^)) 3 , v = 0 on r 0 } , a(u,v)= L(»)= f JTI
fffyOOeyOOdx, fli»,.dr+
f/^idx,
(4.329) (4.330) (4.331)
•In
respectively. Using (4.327), a(u, v) can be written as a(u, v) = I" {AV • uV • v + 2/iEi/n)e£/v)} dx,
(4.332)
from which it is clear that a{-, •) is symmetric. The space V is a Hilbert space for the (H 1 (n)) 3 -norm, and the functionals a(-, •) and L are clearly continuous over V. Proving the F-ellipticity of a(-, •) is nontrivial; actually this ellipticity property follows from the so-called Korn inequality for which we refer to Duvaut and Lions [1] (see also Ciarlet [3] and Nitsche [1]). Now consider the variational problem associated with V, a, and L, i.e.: Find u e V such that a(u,v) = L(v), VveF; (4.333) from the above properties of V,a(-, •), L, the variational problem (4.333) has a unique solution (from Theorem 2.1 of Sec. 2.3). Applying the GreenOstrogradsky formula, (4.33) shows that the boundary-value problem corresponding to (4.333) is precisely (4.328).
394
App. I A Brief Introduction to Linear Variational Problems
The finite element solution of (4.328), via (4.333), is discussed in great detail in e.g., Ciarlet [ l ] - [ 3 ] , Zienkiewicz [1], and Bathe and Wilson [1]. Remark 4.25. The term a(u, v) can be interpreted as the work of the internal elastic forces and L(v) as the work of the external (body and surface) forces. Thus, the equation a(u, v) = L(v),
Vv e V
is a reformulation of the virtual work theorem. 4.7.3. A thin plate problem We follow the presentation of Ciarlet [3, Chapter 1]. Let Q be a bounded domain of R2 and consider V,a(-, •), L defined by V = Hl(il) = Wii)HHa)
= \v\ve H2(Q), v = ~ = 0 on r k (4.334)
8 V a{u, v)= [ \AU\V + (1 -
82u 82v ox 1 ox2 =
d2u 32v\} dx2 ox\))
ioAuAv + (1 - ff)
Jo I.
1
'dh^d2^ \dx2 dx2 ' dx\ dx
ox2 oxl ox2
L(v) = \ fvdx,
feL2(Q).
(4.336)
The associated variational problem: Find ue V such that a(u, v) = L(v), \/veV, (4.337) corresponds to the variational formulation of the clamped plate problem, which concerns the equilibrium position of a plate of constant (and very small) thickness under the action of a transverse force whose density is proportional t o / . The constant a is the Poisson coefficient of the plate material (0 < a < j). If f = 0, the plate is in the plane of coordinates {x1; x2}. The condition u e HQ(Q) takes into account the fact that the plate is clamped. The derivation of (4.337) from (4.333) is discussed in Ciarlet and Destuynder [1]. The Poisson coefficient a satisfies 0 < a < ^; the bilinear form a(-, •) is Ho(Q)-elliptic, since we have a(v, v) = <7||Ac||£,(n) + (1 -
nas
been defined in Sec. 4.5.3.2).
(4.338)
4 Application to the Solution of Elliptic Problems for Partial Differential Operators
395
Actually, using the fact that Hl(Q) = @(Q)H2(n), we can easily prove that a(v,w)= [\oAwdx,
Vv,weH20(Q).
(4.339)
From (4.339) it follows that the solution u of the variational problem (4.337) is also the unique solution in H 2 (Q) of the biharmonic problem A2u = / in Q,
u = ^ = 0, (4.340) on and conversely (here A2 = AA). Problem (4.340) also plays an important role in the analysis of incompressible fluid flows (see, e.g., Girault and Raviart [1], Glowinski and Pironneau [1], and Glowinski, Keller, and Reinhart [1] for further details). For the finite element solution of (4.337), (4.340), see Strang and Fix [1], Ciarlet and Raviart [1], Ciarlet [3], and Brezzi [1], and the references therein. 4.7.4. The Stokes problem As mentioned in Chapter VI, Sec. 5.2, the motion of an incompressible viscous fluid is modelled by the Navier-Stokes equations; if we neglect the nonlinear term (u • V)u, the steady case reduces to the steady Stokes problem - A u + Vp = f i n Q ,
V-u = 0 i n Q ,
u | r = g with
g - n d F = 0; (4.341)
in (4.341), Q is the flow domain (Q c MN, N = 2 or 3 in practical applications), F is its boundary, u = {M;}f=1 is the flow velocity, p is the pressure, and f is the density of external forces. For simplicity we suppose that g = 0 on F (for the case g # 0, see Chapter VII, Sec. 5, and also Appendix III) and also that Q is bounded. There are many possible variational formulations of the Stokes problem (4.341), and some of them are described in Chapter VII, Sec. 5; we shall concentrate on one of them, obtained as follows: Let
• ={&}?=! e ( 0 W (=>c|»|r = 0); taking the [R^-scalar product of <J> with both sides of the first equation (4.341) and integrating over Q, we obtain -
f Au • <> | dx + \\p-^dxJn
JQ
f f • <> | dx Jn
(4.342)
396
App. I A Brief Introduction to Linear Variational Problems
Using the Green-Ostrogradsky formula (4.33), it follows from (4.342) that
(\u-\$dx-
f ~-($>dr- (p\-$dx + [p
Jn
Jron
Jn
Jr
(4.343) (with Vu V(|) = YA=I reduces to
VM< • V^i). Since <$> = 0 on T, the above relation (4.343)
f Vu • V<|> dx = f f • * dx + f pV • <> dx. •Jn
Jn
N o w suppose that §e"V,
(4.344)
Jn
where
TT = {<|>|c|>G(^(Q)r, V-ct> = 0 } ; from (4.344) it follows that f Vu • V<|) dx = f f • <J>
rfx,
V <> | e "T.
(4.345)
Relation (4.345) suggests the introduction of F, a(•, -),L denned by V = {v | v € (tf J(fi))", V • v = 0}, a(v, w) = f Vv • Vw dx, •la
(4.346)
V v, w e (H\Q))N,
(4.347)
L(v)= f f - v d x ,
(4.348)
respectively, and then, in turn, the following variational problem: Find u e V such that a(u, v) = L(v),
VveF.
(4.349) N
Since the mapping v -> V • v is linear and continuous from (Hl(Q)) into L2(Q), F is a closed subspace of (HQ(Q))N and, since Q is bounded is therefore a Hilbert space for the scalar product {v, w} -» j n Vv • Vw dx. The bilinear form a( •, •) is clearly continuous and (Ho(n))iV-elliptic, and the linear functional L is continuous over F if f E (L2(Q))iV. From the above properties of V,a(-. •), L, it follows from Theorem 2.1 of Sec. 2.3 that the variational problem (4.349) has a unique solution. Since we have (cf. Ladyshenskaya [1]) ^(Hi(«))w _
y
it follows from (4.345), (4.349) that if {u, p} is a solution of the Stokes problem (4.341), then u is also the solution of (4.349). Actually the reciprocal property is true, but proving it is nontrivial, particularly obtaining a pressure p e L 2 (Q) from the variational formulation (4.349); for the reciprocal property we refer
5 Further Comments: Conclusion
397
to, e.g., Ladyshenskaya, loc. cit., Lions [1], Temam [1], Tartar [1], and Girault and Raviart [1]. We refer to Chapter VII, Sec. 5 for finite element approximations of the Stokes problem (4.341) (see also the references therein) and also to Appendix III for some complements.
5. Further Comments: Conclusion Variational methods provide powerful and flexible tools for solving a large variety of boundary-value problems for partial differential operators. The various examples discussed in this appendix are all classical (or almost classical) boundary-value problems for elliptic operators, but in fact the variational approach can also be used to solve first-order systems as shown in Friedrichs [1] and Lesaint [1]. As a last example showing the flexibility of variational formulations and methods, we would like to discuss the approximate calculation of the flux associated with the solution of an elliptic boundary-value problem. For simplicity we consider the Dirichlet problem (with Q. bounded): -V-(XVM)
=/inQ,
u = gonT,
(5.1)
whose variational formulation is given (see Sec. 4.3.2) by: Find u £ HX(Q) such that u = g onT and
f (K\u) -\vdx = f fvdx,
V v e Hi(Q).
(5.2)
If the hypotheses on A, /, g made in Sec. 4.3 hold, we know (from Sec. 4.3.2) that (5.1), (5.2) has a unique solution in i/^Q). We call flux the boundary function29
where n still denotes the unit vector of the outward normal at F. There are many situations in which it is important to know (at least approximately) the flux X; this can be achieved through a finite element approximation of (5.1), (5.2) as discussed below. With Hi and H^h as in Sec. 4.5.2, we approximate (5.1), (5.2) by: Find uh e Hi such that uh = ghonT and
dx= f fvhdx, Jn 'Usually X
VvheHlOh,
(5.3)
398
App. I A Brief Introduction to Linear Variational Problems
where gh is an approximation of g belonging to the space yHj, defined by yH{
= {nh\nh e C°(F), 3 vh e H\ such that fih = vh\r}
(actually yHl is also the space of those functions continuous over F and piecewise linear on the edges of 3~h supported by F). Concerning the approximation of X, a naive method would be to define it by Xh = ((K\uh)-n)\r,
(5.4)
which is possible since \uh is piecewise constant over Q; actually (5.4) yields very inaccurate results. A much better approximation of X is obtained as follows. From the Green-Ostrogradsky formula (4.33) we know that X and u satisfy
[ XvdT = f (K\u) • nv dT = f V • (K\u)v dx + f (XV«) • Vi; dx, V v e H\Q). Since — V • (AVu) = / , we finally have
f Xv dT = - f fv dx + f (K\u) - \ v d x ,
Jr
Jn
Jn
VUG H\Q).
(5.5)
Starting from (5.5) to approximate X, we shall define an approximation Xh of X as the solution of the linear variational problem (in which nh is known from a previous computation): Find Xh e yHl such that
[ x h v h d T = - f fvh dx + f (K\uh) • \ v h dx,
V v h e H\; (5.6)
it is easy to see that (5.6) is equivalent to: Find Xh e yHl such that
f V , dr= - [f(ih dx + f (K\uh) • \nh dx,
Jr
Jn
Jn
Vft e yHl,
(5.7)
where p.h is the extension of \xh over Q such that /i,|j. = 0 , V r e ^ , such that 5T n r = 0 . The variational problem (5.7) is equivalent to a linear system whose matrix is symmetric, positive definite, and sparse. As a final comment, we would like to mention that the above method, founded on the application of the Green-Ostrogradsky formula, can also be applied to the computation of fluxes through lines (or surfaces if Q <= R3) inside Q; this is done, for example, in some solution methods for partial differential equations using domain decomposition.
APPENDIX II
A Finite Element Method with Upwinding for Second-Order Problems with Large First-Order Terms
1. Introduction Upwinding finite element schemes have been a subject of very active research in recent years; in this direction we shall mention, among others, Lesaint [2], Tabata [1], Heinrich, Huyakorn, Zienkiewicz, and Mitchell [1], Christie and Mitchell [1], Ramakrishnan [1], Brooks and Hugues [1] and also Fortin and Thomasset [1], Girault and Raviart [2], Johnson [2], Bredif [1], and Thomasset [1], these last five references being concerned more particularly with the Navier-Stokes equations for incompressible viscous fluids. In this appendix we would like to describe a method (due to Bristeau and Glowinski) which can be viewed as an extension of the method introduced by Tabata, loc. cit.; this method will be described in relation to a particular simple model problem, but generalizations to more complicated problems are quite obvious.
2. The Model Problem Let Q be a bounded domain of U2 and Y = dQ. We consider the problem (with e > 0) -eAu + p • Su = / i n Q,
u = 0, on Y,
(2.1)
where p = {cos 9, sin 8}. We are mainly interested in solving (2.1) for small values of e; in the following we shall suppose that / e L2(Q), and we shall use the notation
Problem (2.1) has as variational formulation (see Appendix I, Sec. 4.3): Find u e i/J(Q) such that
e f \u • \v dx + f ^ v dx = f fvdx, Jn
Jn OP
Jn
VueHj(Q),
(2.3)
400
App. II A Finite Element Method with Upwinding for Second-Order Problems
from which we can easily prove, using the Lax-Miligram theorem (see Appendix I, Sees. 2.3 and 4.3), the existence of a unique solution of (2.1) in
EXERCISE 2.1.
Prove that (2.3) has a unique solution.
Hint: Use the fact that
f ^vdx = 0, JadP
Vve Hl(Q).
3. A Centered Finite Element Approximation We suppose that Q is a bounded polygonal domain of R2. Let {3~h}h be a family of triangulations of Q like those in Chapter VII, Sec. 5.3.3.1; to approximate (2.1), (2.3), we use the space H'oh = {vh\vh e C°(Q), vhlT e Plt V T e ^ , vhlr = 0}.
The obvious approximation of (2.3) using HQH is: Find uh e fljft such that, V vh e Hlh, e
Vuh-Vvhdx+ Jn
^£vhdx= Jn op
fvhdx.
(3.1)
Jn
Problem (3.1) has a unique solution; moreover, if {^h}h is such that the angles of ^ are bounded from below by 60 > 0, independent of h, we have lim \\uh - w||H>(n) = 0. If £ is "too small," it is well known that uh obtained from (3.1) is afflicted with spurious oscillations (unless h is very small); it is therefore necessary to use approximations of (2.1), (2.3) more sophisticated than (3.1).
4. A Finite Element Approximation with Upwinding We have in mind an approximation of (2.1), (2.3) such that e\Vuh-Vvhdx Jn
+ (-^,vh) = \fvhdx, \ op Jh Jn
MvheHlh,
uheHloh,
(4.1)
where the scalar product (•, )ft is defined by («*, »*)» = \
I meas(T) £ uh(MiT)vh(MiT),
where MiT, i = 1, 2, 3 are the vertices of T.
(4.2)
4 A Finite Element Approximation with Upwinding
401
Figure 4.1
M,
We look for an approximation dhuh/dp of du/dfi backward with respect to P, second-order accurate (unlike the approximation in Tabata [1] which is only first-order accurate), and defined at the vertices of 2Th. To define such an approximation, we shall use an interpolation method defined as follows. Let M; be a node of STh such that Mt i F.
(4.3)
Let S; be the half-line starting at Mf and directed by - p (see Fig. 4.1). To express (dhnh/dP)(Mj), we look for two points of S; (different from M;) also belonging to the edges of triangles close to Mt. Let us now describe the choice of these two points. Step 1. Among the triangles of 2Th having Mt as a vertex, we consider the triangle denoted by Jn which is crossed by St. Let Mjt Mk be the two other vertices of Jn (taken with the trigonometric orientation); since (4.3) holds, Jn always exists. Step 2. Denote by Mn the point at the intersection of S( and of the side MjMk. If MjMk is supported by the boundary F, then go to Remark 4.1; if the contrary holds, let Ji2 be the triangle adjacent to Jn along MjMk (see Fig. 4.1) and let M, be the third vertex of J ;2 ; finally, denote by Mi2 the point where St crosses MjML or MtMk. Step3. We have constructed three points on St, which are Mt, MiU Mi2. But M a , Mi2 may be very close to each other and even coincide with either Mj or Mk; if this unfortunate situation holds, our second-order approximation degenerates into a first-order one. Thus we may need to define a different Mi2 if Syis "close" to one of the eges of Jn. We shall proceed as follows. We compute the two angles (MtMj, — P) and (MtMj, MtMk) and then consider the following possibilities (other strategies are possible): (a) ifUMtMj, MtMk) < (MiMj, ~P) < then Mi2 is defined as in Step 2.
402
App. II A Finite Element Method with Upwinding for Second-Order Problems Mt
Figure 4.2
(b) If (MJ^MJ, - P ) < ftMiMj, MiMk), denote Mv = My, if (MtMj, - P ) > UMiMj, MiMk), denote Mv = Mk. If Mi2 of Step 2 does not belong to MvMh then Mn, Mi2 will not coincide, and we can use the Mi2 defined in Step 2 (see Fig. 4.2 (where Mv = Mj)). If Mi2 belongs to MvMl, then |M n Mi2 | may be "very small," and we have to modify the definition of Mi2; go to Step 4 for such a modification. Step 4. We look for the triangle Ji3 adjacent to Ji2 along MVM,. If MVM, is supported by F, such a triangle Ji3 does not exist; then go to Remark 4.1. If Ji3 exists, let Mm be the third vertex of Ji3 (see Fig. 4.3); if S ; crosses edge M ; M m , we denote by Mi2 the crossing point, and one uses this point (together with M^ Mn) to approximate du/dfi; if St crosses MvMm, we repeat the process of this step to finally obtain the following situation: either we have to go to Remark 4.1 1 or we obtain Mi2 as the intersection of St and of the polygonal line, the boundary of the polygonal domain consisting of the union of those triangles of 9~h with Mv as a common vertex. If we have been able to avoid ending at Remark 4.1, we have constructed two points Mn, Mi2 "close" to vertex Mt. Since uh is continuous on fi, its values 1
This cannot happen if Mv $ F.
4 A Finite Element Approximation with Upwinding
403
Figure 4.3
M
at M ; , Mn, Mi2 are known; we denote wjo = uh(Mt),
un = uh(Mn),
(4.4)
ii2 = uh(Mi2).
Let hn = |M t M n |, hi2 = \ MiMi2 |; we then define (dhuh/dP)(Mi) as the value at M ; of the derivative of a second-order polynomial, defined on the half-line S ; and coinciding with uh at M ; , Mn, Mi2. We then obtain
(M ; ) = ^ j - ± ^ ui0 hnhi2
fe
"
M|1 +
hn(hi2 - /i(1)
fe
"
« i 2 . (4.5)
hi2(hi2 - hn)
Remark 4.1. If Mt is close to F and if we have not been able to define Mi2, we can proceed as follows: (i) Either define accurate—by
a centered
approximation
of
(du/dP)(Mi)—O(h2)
(4.6) where !fi is the subset of 9~h consisting of those triangles with M, as a common vertex; At — meas((J r s ; r . T); duJdP\T = P • Vuh\T (we recall that \uh is piecewise constant on Q).
404
App. II A Finite Element Method with Upwinding for Second-Order Problems
(ii) Or use an upwinded first-order approximation (as in Tabata [1]) defined by
dp
(4.7) Ju
where the triangle Jn has been defined in Step 1.
5. On the Solution of the Linear System Obtained by Upwinding Using the upwinded scheme described in Sec. 4, we obtain uh via the solution of a linear system whose matrix has a bandwidth which is approximately twice the bandwidth of the matrix associated with the centered approximate problem (3.1) of Sec. 3. The matrix of the above system is definitely nonsymmetric; in the preliminary stage of our numerical experiments, we have chosen to solve the linear system, say AUft = b,,
(5.1)
by the standard method consisting of solving the normal equation A ^ U , = MhK,
(5.2)
by a Cholesky method, since the matrix in (5.2) is symmetric positive definite. At the moment, we are testing more sophisticated methods like the Lanczostype methods described in, e.g., Widlund [1] (see also Strikwerda [1] for iterative methods for solving finite difference approximations of second-order elliptic equations with large first-order terms).
6. Numerical Experiments 6.1. The test problem
We have chosen to solve -sAu + P- VM = 1 inQ, u = 0 on T,
(6.1)
where Q is the domain (cavity with a step) of Fig. 6.1. 6.2. Numerical results The method of Sec. 3, using a centered approximation, produces very poor results as soon as e < 10""l (at least with the finite element grids shown in Figs 6.2-6.5); therefore we only present results concerning the upwinding
6 Numerical Experiments
405
X,'
fi
Figure 6.1
method of Sec. 4. For the nodes close to F, we have used the first-order approximation defined by (4.7). The domain Q being the one of Fig. 6.1, we have used the regular triangulations of Figs. 6.4 and 6.5 in our tests. We have then solved (6.1) via the finite element approximate problem (4.1). The above tests are severe since the solution contains strong boundary layer and free layer (generated by the corner) phenomenon, resulting in strong variations of u and/or VM.
Figure 6.2. Triangulation ST\ (360 triangles, 211 nodes).
406
App. II A Finite Element Method with Upwinding for Second-Order Problems
Figure 6.3. Triangulation 9~\ (1000 triangles, 551 nodes).
Figure 6.4. Triangulation 9~\ (472 triangles, 267 nodes).
6 Numerical Experiments
Figure 6.5. Triangulation ST\ (1246 triangles, 674 nodes).
Figure 6.6. Triangulation 9~\ (s = 10" 3, 9 = 0).
407
408
App. II A Finite Element Method with Upwinding for Second-Order Problems
Figure 6.7. Triangulation 3~l (s = 10" 3, Q = 0).
Figure 6.8. Triangulation 9~\ (s = 10~3, d = 0).
6 Numerical Experiments
409
Figure 6.9. Triangulation ,Tt {& = 10 " 3 , 9 = 0).
Figure 6.10. Triangulation $~l{e=
10" 3 , 0 = re/3).
410
App. II A Finite Element Method with Upwinding for Second-Order Problems
Figure 6.11. Triangulation
Figure 6.12. Triangulation S~\ (e = 10 " 3 , 6 = rc/3).
6 Numerical Experiments
Figure 6.13. Triangulation 3~* (e = 1(T 3 , 8 = rc/3).
Figure 6.14. Triangulation 9~\ (e = 10" 5 , 6 = 0).
411
412
App. II A Finite Element Method with Upwinding for Second-Order Problems
Figure 6.15. Triangulation 9~i (e = 10"5, 9 = 0).
Figure 6.16. Triangulation y\ (s = 10~5, 9 = jt/3).
6 Numerical Experiments
413
Figure 6.17. Triangulation $~% (?, = 10" 5 , 8 = TI/3).
Using the notation p = {cos 9, sin &}, we show on Figs. 6.6-6.17 the equipotential lines of the approximate solutions corresponding to various values of e and 6. The agreement between these results is good, except if 6 = 0, for which the results obtained with the regular triangulations 3~\, 9~\ are good, unlike those obtained with 3~\, 2T\. In fact, the case 6 = 0 is quite severe, since it is a case for which the limit, as e -• 0, of the solution of (6.1) is a discontinuous function; indeed it is easily shown that this limit is u(xu x2) = Xi
if 1 < x2 < 2,
(6.2)
u(xu x2) = x x — 1 if 0 < x2 < 1, and is therefore discontinuous along the line x2 = 1, with a jump of amplitude 1; actually the numerical results obtained with ST\, 9~\ for 6 = 0 and very small e are nearly optimal, since the discontinuities of the limit function (6.2) (including the boundary layers) are supported only by one layer of triangles on which the approximate solution varies continuously without oscillations. The above results can be improved by including special devices to take into account the behavior of the solution in the boundary layers. The poor quality of the results obtained with &~\, 2T\, for 6 — 0 and e very small, seems to be due to the fact, pointed out by Engquist and Kreiss [1], that irregular meshes may introduce very severe phase distorsions. For 0 / 0 (and 6 / n/2), the limit solution as e -> 0 is continuous (but not C 1 ), and therefore the numerical process, is less sensitive to phase distorsion (but exhibits some numerical diffusion (as shown on Figs. 6.10-6.13, 6.16, 6.17) along the line starting from
414
App. II A Finite Element Method with Upwinding for Second-Order Problems
the corner and propagating the discontinuities of the derivatives of the limit function).
7. Concluding Comments The finite element scheme with upwinding that we have introduced in Sec. 4 of this appendix is quite accurate and extremely robust; it may also handle fairly irregular geometries. The above scheme leads, of course, to a nontrivial coding. We are presently extending the ideas of Sec. 4 to the solution of more complicated problems (in fluid mechanics, semiconductor simulations, etc.).
APPENDIX III
Some Complements onthe Navier-Stokes Equations and Their Numerical Treatment
1. Introduction In this appendix, wou would like to complete Sec. 5 of Chapter VII, where the numerical solution of the Navier-Stokes equations for incompressible viscous fluids was discussed. In Sec. 2 we shall discuss the finite element approximation of the Dirichlet boundary condition u = g on T, with | r g • n dT = 0, when g # 0. In Sec. 3 we shall make comments on the treatment of the nonlinear term, (u • V)u. In Sec. 4, boundary conditions other than the Dirichlet boundary condition, will be discussed, and their effect on the decomposition properties of the Stokes problem will be considered in Sec. 5.
2. Finite Element Approximation of the Boundary Condition u = g on F if g & 0 Let us consider the steady-state Navier-Stokes equations, i.e., -vAu + (u- V)u + \p = f inQ, u = g on F, with
V-u = 0infi,
g • n dT = 0
(2.1)
(n is the unit vector of the outward normal). In Chapter VII, Sec. 5.3.3, we have defined the following finite element spaces: Vh = {v,|vft e C°(Q) x C°(Q), v h|T e P2 x P2, Vgh={yh\yheVh,yh = g,onr},
VTe3Th},
(2.2) (2.3)
where gh is a convenient approximation of g (the following also holds if Vh is defined by (5.21) of Chapter VII, Sec. 5.3.3.1).
416
App. Ill Some Complements on the Navier-Stokes Equations
A problem of important theoretical and practical interest is to construct gh such that
1
gh-ndT = 0;
for simplicity we suppose that g is continuous over P. We now define the space yVh as V*i={mll»» = v*lr,
vfieFJ,
(2.4)
i.e., yVh is the space of the traces on T of those functions \h belonging to Vh. Actually, if Vh is defined by (2.2), yVh is also the space of those functions defined over F, taking their values in U2, continuous over T and piecewise quadratic over the edges of 2Th contained into P. Our problem is to construct an approximation gh of g such that h,
\gh-ndT
= O.
(2.5)
Jr
If nhg is the unique element of yVh, obtained from the values taken by g at those nodes of 2Th belonging to T, we usually have j r nhg • n dT # 0. To overcome the above difficulty, we may proceed as follows: (i) We, define an approximation nh of n as the solution of the following linear variational problem in y Vh: h,
\nh/ihdr = \ n/thdr, (2.6) Jr Jr problem (2.6) is in fact equivalent to a linear system whose matrix is sparse, symmetric, positive definite, and quite easy to compute, (ii) Then define gh by -
7
jp K.
(2.7)
It is easy to check that (2.6), (2.7) imply (2.5).
3. Some Comments On the Numerical Treatment of the Nonlinear Term (« • V)« Let us denote by B(a) the nonlinear term (u • V)u occurring in the NavierStokes equations (see Chapter VII, Sec. 5.2). An important property of operator Bis
1 n
B(v)-vdx
= 0,
V v e (Hj(Q))N such that V • v = 0.
(3.1)
4 Further Comments on the Boundary Conditions
417
If V • v # 0, we have JnB(v) • v dx # 0, in general. If we now consider the various alternating-direction methods discussed in Chapter VII, Sec. 5, we observe that we do not require the incompressibility condition V • u = 0 to be satisfied for the nonlinear steps. In order to improve the well-posedness properties of the elliptic problems to solve at these nonlinear steps (and also to simplify convergence proofs), one may replace the original nonlinear term B(u) by = (u • V)u + iu(V • u), following Temam [1]. It is clear that B(n) = B(u) if V • u = 0. Actually the good property of B is that f 5(v) • v dx = 0,
V v e (H^G))" (even if V • v # 0)
(3.2)
(see Temam [1] for further details). 3.1. Prove (3.2) A similar observation holds for the approximate problems discussed in Chapter VII, Sec. 5. However, we have to mention that the numerical results obtained using either B or B are practically identical, provided that At is sufficiently small (At is the time step occurring in the alternating-direction methods of Chapter VII, Sec. 5).
EXERCISE
4. Further Comments on the Boundary Conditions 4.1. Synopsis In Chapter VII, Sec. 5, we have considered the solution of the Navier-Stokes equations for boundary conditions of Dirichlet type. Actually, in many problems, these Dirichlet boundary conditions are inappropriate; they have, for example, the property of being too reflecting if one approximates flows in an unbounded region of UN (like flows around an obstacle or in a channel) by flows in a bounded domain. The goal of this section is to show that the methods discussed in Chapter VII, Sec. 5 can be generalized to the solution of the following Navier-Stokes problem: au - vAu + (u • V)u + Vp = f in Q, V • u = 0 in Q, u = g0onr0,
v — -pn = g1onT1.
(4.1)
418
App. Ill Some Complements on the Navier-Stokes Equations
In (4.1), v, u, p, f are as in Chapter VII, Sec. 5.2, ot is a non-negative constant (a = 0 for a steady flow, a > 0 if (4.1) follows from the time discretization of the time-dependent Navier-Stokes equations), n is the unit vector of the outward normal at F; finally g 0 , g1 are two functions defined over F o , F l 5 respectively, where F o , F x are two subsets of the boundary F of Q such that
\dr>o,
Vi = o,i,
r o u r , = r,
f
dr = o; (4.2)
a partition of F obeying (4.2) is shown on Fig. 4.1. Several continuous and discrete variational formulations of (4.1) will be discussed in Sees. 4.2 and 4.3, and their decomposition properties will be discussed in Sec. 5. Some (useful) generalizations of the results of this section may be found in Conca [1]. Remark 4.1. Suppose } F l dY > 0; then the boundary condition v
3n "a
n
P = gl
o n
1l
on adjusts the pressure without ambiguity (unlike the Dirichlet boundary condition u | r = g for which the pressure is determined only to within an arbitrary constant).
4.2. Variational formulations of (4.1). (I) The continuous case For simplicity we suppose that f e (L2(Q))JV in the sequel, and also that Q is bounded. 4.2.1. Standard variational formulations Let us define the space J f 0 by f, v = 0 on F o }.
(4.3)
Now multiplying the first relation (4.1) by v e Jt0, and integrating by parts (i.e., applying the Green-Ostrogradsky formula (4.33) of Appendix I, Sec. 4), we obtain a
u • v dx + v Vu • Vv dx + Jn Jn
JQ
(u • V)u • v dx — \ p\ • v dx Ja
4 Further Comments on the Boundary Conditions
419
Figure 4.1
which combined with the last relation (4.1), implies, that any pair {u, p}, a solution of (4.1), has to satisfy a \ u-v dx + v \ \u-\\ Jn Jn
= [f-ydx
dx +
Jn
+ f gi --\dT,
\ pV • \ dx Jn (4.4)
Vve
Jrx
Jn
(u • V)u • v dx -
We now suppose that g 0 is sufficiently smooth to be the trace on T o of a function g 0 e (H1(Q))N (if I \ is a "good" subset of F—which is always the case in practice—-we can choose g 0 to have V • g 0 = 0 on Q); it is then quite natural—from (4.4)—to introduce the following variational problem: Find {u, p} e (// 1 (^)) iV x L2(O) such that u = g 0 on T o and a
u • v dx + v Jo
\u • Vv dx + Jn
= f f -\dx Jn
Jn
+ f gj -\dT, Jr,
(u • V)u • v dx — \ p\ • v dx Jet VveJf0,
V-u = 0mQ. (4.5)
If we now restrict the test functions v in (4.5) to the space Vo defined by Vo =
(4.6)
formulation (4.5) yields the variational problem: Find {a, p} e (H^Q))^ x L 2 (Q) such that u = g0 on T o , V • u = 0 in Q, and
J
n
u • v dx + v = [i-ydx Jn
Jn
Vu • Vv dx +
+ f gi-vdr, Jr,
Jn
(u • V)u • v dx VveF0>
(4.7)
in which the pressure p is (apparently) eliminated. Proving the existence of solutions for (4.5), (4.7), and also the equivalence between these problems and (4.1), is beyond the scope of this book, and therefore will not be discussed here (indeed these existence and equivalence results hold if we suppose that gx e (L2(F1))IV, for example). However, proving the
420
App. Ill Some Complements on the Navier-Stokes Equations
existence and uniqueness of solutions for the corresponding Stokes problem (i.e., for the problem obtained by neglecting the nonlinear term (u • V)u) is not too difficult and left as an exercise to the reader. 4.2.2. A less standard variational formulation Let {u, p} be a solution of (4.1); multiplying the first relation (4.1) by v + \(j>, where v e / 0 and (j) e 3>(Q), and since V • u = 0 on Q, we obtain r a
r u • v dx + v
Jn
-
r Jn
r
Vu • Vv dx + (u • V)u • (v + V>) dx Jn Jn
p(A(p + V • v) dx
f =
f f-(v + \6)dx +
Jn
gi-vdr,
V{v,
Jr,
(4.8) From the density property &(Q)"2 = tfo(Q),1 relation (4.8) also holds for any {v, ^ } e j ^ x Hl(Q). We now define two spaces, Wo and Wg, by
f
,
f
Wo = < {v, 6} I iv, 6} e Jf0 x Hk(Q.\
{ = [\-\wdx, Vweff'pl, Jn J
Vci • Vw dx
Jn (4.9)
W, = \ {v, >} I !v, 6} e (H^Q))^ x Hk(Q.), v = g 0 on F o ,
{ r
=
Jn
1 V • vw dx, Mwe H\Si)},
Jn
S
(4.10)
J
respectively. A key result is given by: Lemma 4.1. Suppose that the boundary T of Q is sufficiently smooth (or Q. is convex). Then, if {v, <$>} e Wo or Wg, we have ~Acj) = V- v inQ, (j) e H&Q).
(4.11) (4.12)
PROOF. Since Wo is the particular case of Wg, corresponding to g0 = 0, we prove the lemma for this latter space. Consider {y, cf>} e Wg; we then have (since HQ(Q) a j
= {(f>\(f> 6 H 2 (Q), ^> = d(j)/dn = 0 on F}.
4 Further Comments on the Boundary Conditions
421
and (* \(j) • Vw dx = I* V • vw dx,
V w e H^Q.),
which implies that <> / is the unique solution in Hj(fi) of the Dirichlet problem -A(t) = Vvinfl,
0 = Oonr.
(4.13)
2
Since v e (tf'(Q))", we have V • v e L (Q); it then follows from Chapter I, Sec. 2.4 that the smoothness of T (or the convexity of Q) implies the regularity property cj) e H2(Q) n Hj(fi).
(4.14) 1
Multiplying the two sides of the first relation (4.13) by w e H ^), we obtain, from Green's formula,
- f A
Vwetf 1 ^). (4.15)
But since {v, 0} e H^ implies \V
V w e //H^),
= 0,VH>£
H 1 ^), which in turn implies
- = 0onT; (4.16) on (4.14), (4.16) imply (4.12) and complete the proof of the lemma. • From Lemma 4.1 it is feasible to take {v, >} e Wo in (4.8), and since (from (4.11)) A
x L2(Q) is solution of (4.1), then {u, 0} e Wg.
(4.17) Collecting the above results and properties, we have finally proved that if {u, p} is solution of (4.1) (and (4.5), (4.7)), then {a, 0} e Wg and a
r Jn
u • \ dx + v
r Jn
Vu • Vv dx +
r Jn
(u • V)u • (v + V>) dx
= ff-(v + \(j))dx + f gi-vdr, V{v, 4>}eW0. Jn JrL It is natural to introduce the following variational problem:
(4.18)
Find {u, \j/} e Wg such that a
u • v dx + v Jn
Vu • Vv dx + Jn
(u • V)u • (v + \<j)) dx Jn
= f f -(v + \(f>)dx + f gi - v d r , Jn JTI
V{T,^}eW 0 .
(4.19)
422
App. Ill Some Complements on the Navier-Stokes Equations
We have seen that any solution {u, p} e (Hl(Q))N x L 2 (Q) of (4.1) (and (4.5), (4.7)) is such that {u, \\J} is solution of (4.19) with \jj = 0. Actually the reciprocal property is true and we have: Proposition 4.1. There is equivalence between (4.19) and (4.1); moreover, if {u, i/4 is solution o/(4.19), we have i// = 0. PROOF. We have already proved that if {u, p} e (H1(Q))N x L2(fi) is solution of (4.1) (and therefore of (4.5), (4.7)) then {u, 0} is solution of (4.19). We now prove the reciprocal property. Let {u, \j/} e Wg be a solution of (4.19); from Lemma 4.1 we know that — Ai/> = V • u in Q and that ip e Hl(Q). Since \jj e Hl(Q), we have Vi/f 6 (H J(fi))N and in fact {-\IJ/,\IJ}
eW0.
(4.20)
From (4.20) it follows that we can take {v, 4>} = { — \\j/, <j/} in (4.19); we then obtain -a
u-\i//dx-v
Vu V(Vi/0 dx = a j V • u\j/ dx - v ) Aij/\ • u dx = 0. (4.21)
Then combining (4.21) and — Ai// = V • u, we in turn obtain a f IViAl2 dx + v f |AiA|2 dx = 0,
(4.22)
which implies that \p = 0 and also V • u = — Ai/> = 0. Thus we have proved that if {u, i//} is solution of (4.19), then \j/ = Oandu e Vg. Now consider v e Vo; we clearly have {v, 0} e Wo, and by substitution in (4.19) we finally have proved that u satisfies neVg,
a. u • v dx + v «/Q
Vu • Vv dx 4»*J2
= [ f - v d x + f gj-vdr,
(u • v)u • v dx
•'SJ
VveF 0 ,
i.e., u is a solution of (4.7), which, by equivalence with (4.1), completes the proof of the proposition. Remark 4.2. Proposition 4.1 generalizes Theorem 5.1 of Chapter VII, Sec. 5.3.2.2. 4.3. Variational formulations of (4.1). (II) Finite element approximations We suppose that Q is a bounded polygonal domain of U2. 4.3.1. Triangulation o/Q. Fundamental discrete spaces We follow Chapter VII, Sec. 5.3.3.1. Let {^h}h be a family of triangulations of Q such that Q = UTeS-h T. We set h(T) equal to the length of the greatest side of T, h = max r e 5 - h h(T), and we suppose that the points of T at the interface of T o and Tt are vertices of ^ .
(4.23)
4 Further Comments on the Boundary Conditions
423
We then define the following finite element spaces: Hi = {4>h 6 C°(U),
V T e JTh},
1
Hhh = {>„ e H, , ^ = 0 o n r } = Hi n H = {vh e (C°(Q))2, v,,T e P2 x P 2 , V9h = {*h £K,Vh
(4.24)
tf£(Q),
VT e ^ } ,
(4.25) (4.26)
= go/, o n T o } ,
(4.27)
where gOft is a convenient approximation of g 0 (if g 0 = 0, one takes gOh = 0 in (4.27) to obtain VOh),
Wgh = i{yh, 4>h} € Vgh x Hhh, J ^4>h • VwA dx = J V • vhwh dx, \/wh (4.28) As in Chapter VII, Sec. 5.3.3.1, we can also use the variants of Vgh, Wgh obtained from Vh = K £ (C°(U))2, T ^ e ^ x P , ,
V Te?„},
(4.29)
where ^ is the triangulation of Q obtained from ^ by subdivision of each T e 2Th into four subtriangles, by joining the mid-sides (see Fig. 5.1 of Chapter VII, Sec. 5.3.3.1). 4.3.2. A first approximation of the Navier-Stokes equations (4.1) This is the finite-dimensional analogue of formulation (4.5) defined as follows: Find{ah, ph} e Vgh x Hi such that a \ uh • \h dx + v \ \uh • \\h dx + (uA • \)uh • \h dx Jn Jn Jn -
\ph\-\hdx= Jn (\-uhqhdx Jn
\fh-\hdx+ Jn = 0,
Jr!
glh-\hdT,
V\heVOh,
\/qheHl,
(4.30)
where ih and glh are convenient approximations of f and g1, respectively. Formulation (4.30) generalizes (5.156) of Chapter VII, Sec. 5.8.3. 4.3.3. A second approximation of the Navier-Stokes equations (4.1) This is the finite-dimensional analogue of formulation (4.19) defined as follows:
424
App. HI Some Complements on the Navier-Stokes Equations
Find {nh, \j/h} e Wgh such that r r r a \ uh • \h dx + v Vu,, • Vvh dx + \ (uh • V)uh • (\h + \(j>h) dx Jn Jn Jn r =
r
Mh • (Vft + \
V {\h, 4>h} e WOh;
(4.31)
Jri
Jn
formulation (4.31) generalizes (5.38)! of Chapter VII, Sec. 5.3.4. Actually any pair {uh, il/h}, a solution of (4.31), is characterized by the existence of a discrete pressure ph e H\ such that Sph • \whdx
+ \ (uh • \)uh • \whdx=
a. \ uh • yhdx Jn
+v
V w » e H^h,
VuA • \\h dx + \ (uh • V)uh • \h dx ph\ • yh dx Jn Jn Jn
= \ fh-yhdx+ Jn
\fh- \whdx,
g lfc • yh dT,
V\h e VOh,
{uh, \j/h} e Wgh,
ph eHj,.
Jri
(4.32)
The decomposition properties of the approximate problem (4.31) will be discussed in Sec. 5. 4.4. Further comments The boundary condition v ^ - / , n = glonr1 dn
(4.33)
has no obvious physical meaning. It has, however, the advantage—as mentioned before—of being less reflecting than the Dirichlet boundary condition; also it occurs in a natural way when some domain decomposition techniques are applied to the solution of the Stokes and Navier-Stokes equations. A more physical boundary condition would be to specify the normal stress on r\, i.e., to consider en = g t on r l 5 where a = (crl7) is the stress tensor defined by
ij-. Kronecker's symbol).
(4.34)
5 Decomposition Properties of the Continuous and Discrete Stokes Problems of Sec. 4.
425
As may be expected, the numerical treatment of (4.34) is more complicated than that of (4.33) and will not be discussed here.
5. Decomposition Properties of the Continuous and Discrete Stokes Problems of Sec. 4. Application to Their Numerical Solution 5.1. Synopsis The main goal of this section is to show that the decomposition properties ofthe Stokes problem discussed in Chapter VII, Sec. 5(whenF 0 = F, Tl = 0) still hold for the Stokes problem associated with the various formulations of the Navier-Stokes equations (4.1) discussed in Sec. 4 of this appendix. 5.2. Formulations of the continuous and discrete Stokes problems The notation are those of Sec. 4 of this appendix. The Stokes problem that we consider is au - vAu + \p = f in Q, u = g 0 on T o ,
v
V • u = 0 in Q, pn = gt on r \ .
on A first variational formulation of (5.1) is: Find {u, p} e (H\n)f x L2(Q) such that u = g0 on T o and a I u • v dx + v Vu • Vv dx — pV • v dx Jn Jn Jn
= (f-vdx+
f g!-vdr,
vvei& V • u = 0 in Q. (5.2)
(5.2) is the Stokes version of (4.5) of Sec. 4.2.1. A second variational formulation of (5.1) is: Find {u, 4>} e Wg such that a
a • v dx + v Jn
Vu • Vv dx Jn
= f f • (v + V0) dx + f Jn
gl
• v dT,
JTI
which is the Stokes version of (4.19) of Sec. 4.2.2.
V {v, 4>} e Wo,
(5.3)
426
App. Ill Some Complements on the Navier-Stokes Equations
From Sec. 4.3 it follows that the discrete problems corresponding to (5.2) and (5.3) are: Find {uh, ph] e Vgh x Hi such that a \ uh-\hdx Jn
+ v I \uh- \\h dx - \ phV • \hdx Jn Jn
= \ih-yhdx+ (\-uhqhdx Jn and find {uh, \j/h) e Wgh such that a
Jn
f glh-\hdr, = 0,
Vv A eF Oft ,
VqheHl
(5.4)
uh • \hdx + v \ \uh • \\h dx Jn
= \fh-(yh Jn
+ V0J dx + f g1A • yh dT, Jrt
V K , 4>h} e W0h, (5.5)
respectively. Actually (5.5) is equivalent to: Find {nh, iph} e Wgh and ph e H\ such that Ph-\whdx=
(fh-\whdx, Jn
MwhsHlh,
a \ uh • \hdx + v \ \uh • \vh dx - \ ph V • \h dx Jn Jn Jn =
f h -v A rfx+ Jn
glh-vhdT,
M\heVOh.
(5.6)
JTI
5.3. Solution of (5.1) via (5.2), (5.4) We now consider an algorithm which is a direct generalization of algorithm (5.311), (5.312) of Chapter VII, Sec. 5.8.7.4.3; this algorithm is denned as follows: p° e L2(£l) arbitrarily given; +1
then for n > 0, p" being known, we compute u" and p"
(5.7) by
xa" - vAu" = f - Vp" in Q, u" = g0 on To, du" v — = g1+pnnonru on
p"+1 = p " - p V - u " ,
(5.8)
p>0.
(5.9)
5 Decomposition Properties of the Continuous and Discrete Stokes Problems of Sec. 4.
427
We observe that the boundary condition on Tt is quite formal since p", as an element of L2(Q), usually has no trace on Tl; to overcome this difficulty, we shall use a variational formulation of (5.8), namely une(H\Q))N,
u" = g 0 o n r 0 ,
and
a I u" • v dx + v ( Vu" • Vv dx = \ f • v dx + \ p" V • v dx + \ gx • v dT, •In Jn Jn Jn Jr, V v e (H\n))N,
v = 0 on T o .
(5.8)'
About the convergence of algorithm (5.7)-(5.9), we have: Proposition 5.1. Suppose that 0
—.
(5.10)
We then have, V p° e L2(Q), lim {u", p"} = {u, p} strongly in (H^Q.))" x L2(Q),
(5.11)
* + oo
where {u, p} is the solution of (5.1), (5.2). PROOF. Define u" and p" by u" = u" - u and p" = p" - p. We clearly have u" e (H\ayf, u" = 0 on T o and
a \vT-vdx + v (*Vu"-Vvdx= \f\-ydx, •>si
•'n
M y e{Hl(a))N,
v = 0onr0,
•'n
(5.12) and (since V • u = 0) p»+i = p " - p V - 0 " .
(5.13)
From (5.13) it follows that \\P"\\hm - \\P-+I\\lm
= 2p f P"V • u" dx - p2 f |V • u"|2 •>n
rfx.
(5.14)
•'n
Now taking v = u" in (5.12), and combining with (5.14), we obtain
»\2dx + v f \\u"\2dx)-p2
f|V-u"|2^. (5.15)
Combining (5.15) with relation (5.325) of Chapter VII, Sec. 5.8.7.4.3 (i.e.,
a
f l y l 2 dx +v f I Vv l 2 dx'
vv
428
App. Ill Some Complements on the Navier-Stokes Equations
we finally obtain
which proves the convergence of u" to u, V p° e L2(il), if (5.10) holds (we have to remember that Q bounded implies that
/ r a
\ 1/2 V
r 2
2
\\\ dx + v
\\v\ dx)
is a norm on {v|v e (H1(Q))N, v = 0 on F o }, equivalent to the (H 1(n))N-norm, and this for all a > 0). The proof of the convergence of p" to p is left to the reader (actually we should prove that the convergence of {u", p"} to {u, p} is linear).
Remark 5.1. Using the material of Chapter VII, Sec. 5.8.7.4, it is straightforward to obtain conjugate gradient variants of algorithm (5.7)-(5.9) and also variants derived from an augmented Lagrangian functional reinforcing the incompressibility condition. The same observations hold for the solution of the approximate problem (5.4). Remark 5.2. When using a finite element variant of algorithm (5.7)-(5.9) to solve the approximate problem (5.4), we have to solve, at each iteration, a discrete elliptic system with boundary conditions of the Dirichlet-Neumann type. The solution of such problems has been discussed in Appendix I, Sec. 4. The same observation holds for the conjugate gradient and augmented Lagrangian algorithms mentioned in Remark 5.1 above.
5.4. Solution of (5.1) via (5.3) and (5.5), (5.6) We follow (and generalize) Chapter VII, Sec. 5.7, where the situation F = F o , Fj = 0 was treated. In this section we suppose that j F l dT > 0. The decomposition properties of the Stokes problem (5.1) follow directly from: Proposition 5.2. LetXeH~ 1/2 (F) and let A: H~ 1/2 (F) -+ H1I2(T) be defined by the following cascade of Dirichlet and Dirichlet-Neumann problems. ApA = 0 in Q, - vAu A = — \px in Q,
=
Pi
= X on F,
OonF 0.
(5.16) n ( = An) onTu
V
~dn~
(5-17) -A\iix --= V
• u
i in Q,
= Oon
r,
(5.18)
5 Decomposition Properties of the Continuous and Discrete Stokes Problems of Sec. 4.
429
and then AX=
Then A is an isomorphism from H form a(-, •) defined by
-
1/2
(5.19)
dn
(F) onto Hl/2(T).
a(X, n) = (AX, /i>,
Moreover, the bilinear
V X, pi e H~ 1 / 2 (r)
(5.20)
1/2
(where <•, •> denotes the duality pairing between H (T) and /J~ continuous, symmetric, and H~ 1 / 2 (F)- elliptic.
1/2
(F)) is
We do not give the proof of Proposition 5.2; let us mention, however, that it is founded on the relation
(AXU X2) = a f uAl . u ^ x + v f VuAl • Vu,2 dx, •Jn
V X,, X2 e H~
1/2
(T),
Jn
where uXl, uAz are the solutions of (5.17) corresponding to X = Xl and X = X2, respectively. Application of Proposition 5.2 to the solution of the Stokes problem (5.1). We define p0, u 0 , i//0 as the solutions of, respectively Q,
Ap0 = V- fin <xu0 — vAu 0 = f -- \p0 in Q,
-go
Po =
onr 0 ,
(5.21)
OonF, 3n
8!
+ p 0 n on Fj, (5.22)
- A ^ o = V-
in £2,
"Ao = 0 on F.
(5.23)
The fundamental result is given by: Theorem 5.1. Let {u, p} be the solution of the Stokes problem (5.1). The trace X = p\x is the unique solution of the linear variational equation
V/iEf/-" 2 (r).
(5.24)
If we compare the above theorem to Theorem 5.7 of Chapter VII, Sec. 5.7.1.2, we observe that this time—due to m e a s ^ ) > 0—the trace of the pressure is uniquely denned by (5.24). The same decomposition principles can be applied to the discrete Stokes problem (5.5), (5.6); since the resulting methods are trivial variants of the methods discussed in Chapter VII, Sec. 5.7.2, they will not be discussed here any further, except to say that, again, meas(F 1 ) > 0 implies that the linear system (discrete analogue of (5.24)), providing the trace of the discrete pressure ph, has a unique solution.
430
App. Ill Some Complements on the Navier-Stokes Equations
6. Further Comments The methods, for solving the Navier-Stokes equations, discussed in Chapter VII, Sec. 5, and in this appendix have been generalized by Conca [1], [2], in order to treat a large variety of boundary conditions involving the stress tensor 5 = (ff^)^} defined by a..
= -p3.j + 2vDu(u),
(6.1)
where u = {«;}f= t and D;/u) = ^(duJdXj + 3u/tbc;). In this direction, it is quite convenient to use the following equivalent formulation of the NavierStokes equations:
«« ( -2v£ ^ A » + £ M , g i + |Uy;.infi, j=l OXj
j=i
OXj
i= l
JV,
OXt
(6.2) V-u = 0 inO.
(6.3)
We refer to Conca, loc. cit., for further details (see also Engelman, Sani, and Gresho [1] for the practical finite element implementation of various boundary conditions associated with the Navier-Stokes equations).
Glossary of Symbols
The number on the right indicates the page of first occurrence, the most common symbols are not repeated in the following chapters.
Chapter I IR real line, i.e. the set of all real numbers R = » u { + 00} u {-00} V a real Hilbert space (•, •) and ||-|| scalar product and norm of V, respectively V* the dual space of V a(-, •) a bilinear form, continuous and F-elliptic a ellipticity constant of a{-, •) (a > 0) L a continuous and linear functional from V to IR K a closed convex and nonempty subset of V j a convex lower semicontinuous (l.s.c.) and proper functional IK indicator functional of K £?(V, V) the space of the operators from Kto V, linear and continuous A the unique operator of £C(V, V) such that a(v, w) = (Av, w), / PK \\A\\
3 3
,,,„ \\A\\ = sup
, 3
Moll —— \\V\\
the norm of L, i.e.
||L|| = sup
'-^f\\V\\
J'(v)
3
the unique element of V such that L(v) =(l,i)),Vi)eF the projection operator from Fto K in the ||-|| norm the norm of operator A, i.e.
veV-{0)
\\L\\
V v, w e V
1 1 1 1 1 i 1 1 1 1 2 3
the Gateaux derivative of J at v; defined by , „, . . ,. J(v + tw) - J(v) (J(v), w) = hm — — <^o t 00
if such a limit exists
4
456
Glossary of Symbols
b(-,-) P X h {Vh}h {Kfc}h X rh U 0';,} e
a bilinear form, symmetric, continuous and F-elliptic ellipticity constant of b(-, •) (/? > 0) a continuous and linear functional from V to IR a parameter converging to 0 a family of closed subspaces of V a family of closed nonempty subsets of Kwith Kh a Vh,M h a dense subset of K an operator from ^ to 1^ (or Kh) a dense subspace of V a family of convex, l.s.c. and proper functionals a positive parameter; to converge to 0
6 6 7 9 9 9 9 9 12 12 15
jE
penalty functional; defined from j by ye = - j e the duality pairing between V* and V the Gateaux differential of _/,, at v defined by
15
<•, •> 7J(U)
16
,.,,-. J (v + <w) -j' E (f) x
RN A b A' B R(B) Ker(B) I p(B'B) 0,gf UN+
'
if such a limit exists A'-dimensional Euclidean real space an N x N matrix, positive definite an element of W the transpose matrix of A an M x N matrix the range of B the kernel of B the M x M identity matrix the spectral radius of B'B a functional from UN to R the positive part of gt, i.e. g* = sup(0, gt) the positive cone of 1RN, i.e. IR+ = {v e RN\ v = {uJfLi,v t > 0, V j = 1,..., N}
16 18 18 18 18 18 18 19 21 23 23 24 25
a bounded domain (i.e. a bounded connected open set) of R2 closure of Q. the boundary of £1 the generic point of R2
27 27 27 27
Chapter II Q Q. r = dil x = {xl5 x 2 }
_ll
27
!' 8x2) Cm(Q) supp(«)
space of m-times continuously differentiable real-valued functions for which all the derivatives up to order m are continuous in Q support of the function v = {x \ v(x) # 0}
27 27
Glossary of Symbols
457
Co(fi) = {v | v e Cm(U), supp(u) is a compact subset of Q} NL.p.n = Zw< m ll^fllL-tfi) f o r «eC m (n) where a = {a,,a 2 }; a,,« 2 are nonnegative integers. |a| = a, + a2 and D" = dM/8xf dxf P L (Q) = {v\v measurable, J n |jj(x)| p dx < + oo} (1 < p < +oo) Wm-P(Q) completion of Cm(f2) in the above norm = {v | v e L'(fi), D«v e Lp(Ci), V a, |a | < m} P
W^- (fi) completion of CJ(H) in the above norm Hm(Q) = Wm'2{O) H'SiQ) = WS'2(n) @(«) = CJCfi) dx = dxt dx2 u| r the trace of v on F HJ(Q) = {v\veH^Q), v|r = 0} = ®(n)" 1
27 27 27 28 28 28 28 28 28 28 28 28 28 28 29
^ a finite element triangulation of SI T the generic triangle of 9~h f interior of T h length of the largest edge of the triangles of 9~h Pk space of polynomials in xu x2 of degree
32 32 32 33 33 33 33 34 35 35 36 36 37 38 40 40 40
meas(fi) = measure of Q H / 2 ' c o (ii)= {v\veC1(Q), all first-order derivatives being Lipschitz continuous . over Q.}
43 44
458 v'=-
Glossary of Symbols 44
dx
r = yjx\ + x\ Vh = {vh\vheC°(U), » k = 0 o n r , vh\TePv V Te,Th} YlTv linear interpolate of v at the three vertices of T hT diameter of T (i.e. the length of the largest edge of T) II,, piecewise linear interpolation operator i? a Lagrangian functional L?(I2) = {q | <j e L"(Q), > 0 a.e.}
45 46 52 52 52 54 54 54
yv dF
trace of v on F superficial measure along the boundary F
56 58
— on
(outward) normal derivative
58
A PA ||y|| H1I2(F)
positive cone of L2(F) = {q\q e L2(F), q > 0 a.e. on F} projection operator from L2(F) to A norm of the trace operator y: Hl(Q.) -> L2(F) space of the traces of the functions of H\Q)
64 66 66 70
L°°(r) = {q\q measurable and bounded on F} A = {n | n e L2(F), | fi(x) \ < 1 a.e. on F} /x viscosity of a Bingham fluid plasticity yield of a Bingham
fluid
L 2 (Q),|q(x)| < la.e.}
70 70 78 78 82 82 84
ZM=
I
91
measure of T
96
two real Hilbert spaces such that V cz H,V = H dual spaces of V and H, respectively scalar products of H and V, respectively norms of H and V, respectively a continuous bilinear form from V x V -* U, elliptic in the following sense: 3 a > 0 and X > 0 such that
98 98 98 98
eh\8x18x2
M(T)
<JA2
Chapter III V, H V*, H* (•, •)» ((•. •)) | • |, || • |! «(•, •)
a(v,v)
+ X \ v \ 2 >tx\\v\\2,
VueF
98
Glossary of Symbols
459
X [0, T] L2(0, T; X)
a Banach space whose norm is denoted by ||-||x 98 a time interval 98 a space of time dependent, X-valued functions, defined as follows: L2(0, T; X) = {x | x(t) 6 X a.e. on ]0, T[, x is a measurable function oft,JJ||x(OllS«fc< +00} 98 / an element of L2(0, T; V*) 98 C°([0, T], X) the space of the Z-valued functions continuous on [0, T ] ; also denoted C°(0, T; X) 99 /i a parameter converging to 0 99 (Vh)h a family of closed subspaces of V, approximating V and H as h -* 0 99 aft(-, •) an approximation of a(-, •), as h -> 0 99 /,, an approximation of/, as /* -> 0 99 At a time discretization step 99 ujj approximate solution at time nAt 99 (•, -)k an approximation of (•, •), as h -> 0 99 (Kh)h a family of closed convex nonempty subsets of Vh, approximating Kash-+0 101 (jh\ a family of convex, proper, l.s.c. functionals defined over Vh, approximating) as h -+ 0 103
Chapter IV 0
a real-valued, continuous, nondecreasing function defined on such that 0(0) = 0 0 the real-valued, C1, convex function defined by (t) = jo 4>{ \)dx j the functional defined by j(v) = j n
110 111 111 114 123 128 128 128 128 129 130 134 134 134 134 134
Chapter V V V*
a real reflexive Banach space dual space of V
140 140
460
Glossary of Symbols
K Jo J1
a nonempty closed convex subset of V a convex differentiate functional from VtoU a convex proper l.s.c. functional from F t o U
140 141 141
Dom(J)= {v\veV,J(v)eU}
141
<•, •) duality pairing between F * and V yl a monotone operator from V to F * J^ a real Hilbert space F = ]~[f=1 F; K, closed convex nonempty subsets of Vt
141 141 152 152 153
K = Y\f=iKt
153
Pj COJ
a projection operator from Vi to K( relaxation parameters
153 155
Chapter VI F, H B F, G
two real topological vector spaces 166 an element of if (F, if) 166 two convex proper l.s.c. functionals, from H to U and F to R, respectively 166 ££ a Lagrangian functional 167 r a nonnegative parameter 168 i?r an augmented Lagrangian functional 168 J the functional from F t o R denned by J(D) = F(Bv) + G(u) 168 dom(;) = {x | x e X, j(x) eR} 168 K(B) = range of B 168
Chapter V I I a mapping from UN to RN with F = {flt..., fN} a nonlinear operator from HQ(Q) to H' x(£i) a time discretization step (k > 0) the H ~l (Q) norm denned by
F T k INI *
II/L = 5 T.E. S p Q.h
sup
l^^
1
a curvilinear abscissa abbreviation for trailing edge a slit of the flow domain the pressure an approximation of Q
Nfe = dim Vh B*=W?*i Xh a set of feasible discrete transonic flow solutions x = Argmin f(y) ifxeC and/(x) < f(y), VyeC M
"e
Mach number
195 198 201
203 207 213 213 214 217 217 217 218 218 218 219 220
Glossary of Symbols
461
M
222 243 245 245 245 245 245
jump of a quantity q along a stream line standard polar coordinates flow velocity pressure viscosity coefficient Reynold number density of external forces a symbolic notation for
{r,0} p V
Re f (u • V)u
245
Hi, Hlh, Vh fundamental finite element spaces 248 Vgh, Wgh finite element spaces taking into account the boundary condition u = g on T 248, 249 iff, a subtriangulation of 9~h 249 k l i . n = (Jfi l Vl Jl 2 dx)1'2 249 lyl
=
llyll
2
2
quotient norm defined by |Mlx/ra = Infceiu \\v + cWx where X is a functional Banach space containing the constant functions k a time discretization step T_ = {x |x e T, g(x) • n(x) < 0} H1I2(T) space of the traces of the functions of H^fi) jell2(T) = {,u|/i6H 1/2 (r), J r A(dr = 0 } H - 1/2
249
\\-\\xia
251 253 259 266 266 266 266 268 271 271 293 294
Appendix 1 C field of the complex numbers z conjugate of the complex number z J"z real part of z T injection operator y a subspace of V such that "T = V (5X0 = Dirac measure at x 0 I N L P o = llflltt-".^)
325 325 325 325 326 344 367
367 ,oo,n = m a x | | 3 S I ! ; | | L « 1 ( n )
367
462
Glossary of Symbols
hT pT c» nh CA x CB a> f g u
length of the largest edge of a triangle T supremum of the diameters of all the disks contained in T notation for the inclusion with continuity of the injection an interpolation operator vector product of CA and CB over-relaxation parameter a density of body forces a density of surface forces displacement vector stress and strain tensors, respectively Lamme coefficients Poisson coefficient
Appendix 2
(see previous chapters)
Appendix 3
(see Chapter VII, Sec. 5, for most notation)
5
normal stress tensor; denned by dui 1 /8u, 2 \dxj
du dx
368 368 368 369 379 384 392 392 392 393 393 394
424 430