Full-Note FPR Partition of Unity P-32 Thm2.7
Full-Note FPR Partition of Unity P-32 Thm2.7
Full-Note FPR Partition of Unity P-32 Thm2.7
by
Baisheng Yan
Department of Mathematics
Michigan State University
[email protected]
Contents
Chapter 1. Preliminaries 1
§1.1. Banach Spaces 1
§1.2. Bounded Linear Operators 5
§1.3. Weak Convergence and Compact Operators 7
§1.4. Spectral Theory for Compact Linear Operators 9
§1.5. Nonlinear Functional Analysis 11
Chapter 2. Sobolev Spaces 23
§2.1. Weak Derivatives and Sobolev Spaces 23
§2.2. Approximations and Extensions 27
§2.3. Sobolev embedding Theorems 39
§2.4. Compactness 45
§2.5. Additional Topics 47
§2.6. Spaces of Functions Involving Time 50
Chapter 3. Second-Order Linear Elliptic Equations 55
§3.1. Differential Equations in Divergence Form 55
§3.2. The Lax-Milgram Theorem 58
§3.3. Energy Estimates and Existence Theory 59
§3.4. Fredholm Alternatives 64
§3.5. Regularity 69
§3.6. Regularity for Linear Systems* 76
Chapter 4. Linear Evolution Equations 97
§4.1. Second-order Parabolic Equations 97
§4.2. Second-order Hyperbolic Equations 111
§4.3. Hyperbolic Systems of First-order Equations 121
§4.4. Semigroup Theory 126
Index 143
iii
Chapter 1
Preliminaries
1
2 1. Preliminaries
Then X is a Banach space with norm k · km,∞ . To prove, for example, the completeness
when m = 0, we let {fn } be a Cauchy sequence in X, i.e., assume for any ε > 0 there is a
number N (ε) such that for all x ∈ Ω
But this means that {fn (x)} is a uniformly Cauchy sequence of bounded continuous func-
tions, and thus converges uniformly to a bounded continuous function f (x). Letting m → ∞
in the above inequality shows that kfn − f km,∞ → 0.
Note that the same proof is valid for the set of bounded continuous scalar-valued func-
tions defined on a nonempty subset of a normed space X.
Example 1.2. Let Ω be a nonempty Lebesgue measurable set in Rn . For p ∈ [1, ∞), we
denote by Lp (Ω) the set of equivalence classes of Lebesgue measurable functions on Ω for
which
Z 1
p
p
kf kp ≡ |f (x)| dx < ∞.
Ω
(Two functions belong to the same equivalence class, i.e., are equivalent, if they differ
only on a set of measure 0.) Let L∞ (Ω) denote the set of equivalence classes of Lebesgue
measurable functions on Ω for which
Then Lp (Ω), 1 ≤ p ≤ ∞, are Banach spaces with norms k · kp . For p ∈ [1, ∞] we write
f ∈ Lploc (Ω) iff f ∈ Lp (K) for each compact set K ⊂ Ω.
For the sake of convenience, we will also consider Lp (Ω) as a set of functions. With this
convention in mind, we can assert that C0 (Ω) ⊂ Lp (Ω). In fact, if p ∈ [1, ∞), then as we shall
show later, C0 (Ω) is dense in Lp (Ω). The space Lp (Ω) is also separable if p ∈ [1, ∞). This
follows easily, when Ω is compact, from the last remark and the Weierstrass approximation
theorem.
Finally we recall that if p, q, r ∈ [1, ∞] with p−1 + q −1 = r−1 , then Hölder’s inequality
implies that if f ∈ Lp (Ω) and g ∈ Lq (Ω), then f g ∈ Lr (Ω) and
kf gkr ≤ kf kp kgkq .
Example 1.3. The Cartesian product X × Y , of two vector spaces X and Y , is itself a
vector space under the following operations of addition and scalar multiplication:
Moreover, under this norm, X × Y becomes a Banach space provided X and Y are Banach
spaces.
4 1. Preliminaries
1.1.4. Hilbert Spaces. Let H be a real vector space. H is said to be an inner product
space if to every pair of vectors x and y in H there corresponds a real-valued function
(x, y), called the inner product of x and y, such that
(a) (x, y) = (y, x) for all x, y ∈ H
(b) (x + y, z) = (x, z) + (y, z) for all x, y, z ∈ H
(c) (λx, y) = λ(x, y) for all x, y ∈ H, λ ∈ R
(d) (x, x) ≥ 0 for all x ∈ H, and (x, x) = 0 if and only if x = 0.
For x ∈ H we set
(1.1) kxk = (x, x)1/2 .
Theorem 1.4. If H is an inner product space, then for all x and y in H, it follows that
(a) |(x, y)| ≤ kxk kyk (Cauchy-Schwarz inequality);
(b) kx + yk ≤ kxk + kyk (Triangle inequality);
(c) kx + yk2 + kx − yk2 = 2(kxk2 + kyk2 ) (Parallelogram law).
Theorem 1.6. Every nonempty closed convex subset S of a Hilbert space H contains a
unique element of minimal norm.
Proof. Let S = {x − y : y ∈ M }. It is easy to see that S is convex and closed. Theorem 1.6
implies that there exists a y ∈ M such that kx − yk ≤ kx − wk for all w ∈ M . Let z = x − y.
For an arbitrary w ∈ M , w 6= 0, let α = (z, w)/kwk2 and note that
kzk2 ≤ kz − αwk2 = kzk2 − |(z, w)/kwk|2
which implies (z, w) = 0. Therefore z ∈ M ⊥ . If x = y 0 + z 0 for some y 0 ∈ M , z 0 ∈ M ⊥ , then
y 0 − y = z − z 0 ∈ M ∩ M ⊥ = {0}, which implies uniqueness.
Proof. It is easy to see that B(X, Y ) is a normed space. To prove completeness, assume
that {Tn } is a Cauchy sequence in B(X, Y ). Since
(1.5) kTn x − Tm xk ≤ kTn − Tm kkxk
we see that, for fixed x ∈ X, {Tn x} is a Cauchy sequence in Y and therefore we can define
a linear operator T by
T x = lim Tn x for all x ∈ X.
n→∞
If ε > 0, then the right side of (1.5) is smaller than εkxk provided that m and n are large
enough. Thus, (letting n → ∞)
kT x − Tm xk ≤ εkxk for all large enough m.
Hence, kT xk ≤ (kTm k + ε)kxk, which shows that T ∈ B(X, Y ). Moreover, kT − Tm k < ε
for all large enough m. Hence, limn→∞ Tn = T .
The following theorems are important in linear functional analysis; see, e.g., [?].
Theorem 1.9. (Banach-Steinhaus) Let X be a Banach space and Y a normed space. If
A ⊂ B(X, Y ) is such that supT ∈A kT xk < ∞ for each fixed x ∈ X, then supT ∈A kT k < ∞.
6 1. Preliminaries
Theorem 1.10. (Bounded Inverse) If X and Y are Banach spaces and if T ∈ B(X, Y )
is one-to-one and onto, then T −1 ∈ B(Y, X).
1.2.2. Dual Spaces and Reflexivity. When X is a (real) normed space, the Banach
space B(X, R) will be called the (normed) dual space of X and will be denoted by X*. El-
ements of X* are called bounded linear functionals or continuous linear functionals
on X. Frequently, we shall use the notation hf, xi to denote the value of f ∈ X* at x ∈ X.
Using this notation we note that |hf, xi| ≤ kf k kxk for all f ∈ X*, x ∈ X.
Example 1.11. Suppose 1 < p, q < ∞ satisfy 1/p + 1/q = 1 and let Ω be a nonempty
Lebesgue measurable set in Rn . Then Lp (Ω)∗ = Lq (Ω). The case of p = ∞ is different. The
dual of L∞ is much larger then L1 .
The dual space X ∗∗ of X ∗ is called the second dual space of X and is again a Banach
space. Note that to each x ∈ X we can associate a unique Fx ∈ X ∗∗ by Fx (f ) = hf, xi for
all f ∈ X ∗ . From Corollary 1.13, one can also show that kFx k = kxk. Thus, the (canonical)
mapping J : X → X**, given by Jx = Fx , is a linear isometry of X onto the subspace
J(X) of X ∗∗ . Since J is one-to-one, we can identify X with J(X).
A Banach space X is called reflexive if its canonical map J is onto X ∗∗ . For example,
all Lp spaces with 1 < p < ∞ are reflexive.
where y = f (z)z. To prove uniqueness, suppose (x, y) = (x, y 0 ) for all x ∈ H. Then in
particular, (y − y 0 , y − y 0 ) = 0, which implies y = y 0 . From the Cauchy-Schwarz inequality
we get |f (x)| ≤ kxkkyk, which yields kf k ≤ kyk. The reverse inequality follows by choosing
x = y in the representation.
Corollary 1.16. H is reflexive.
Proof. For any y ∈ H and all x ∈ H, set f (x) = (T x, y). Then it is easily seen that
f ∈ H ∗ . Hence by the Riesz representation theorem, there exists a unique z ∈ H such that
(T x, y) = (x, z) for all x ∈ H, i.e., D(T ∗ ) = H. Moreover, kT ∗ yk = kzk = kf k ≤ kT kkyk,
i.e., T ∗ ∈ B(H) and kT ∗ k ≤ kT k. The reverse inequality follows easily from kT xk2 =
(T x, T x) = (x, T ∗ T x) ≤ kT xkkT ∗ kkxk.
Proof. To prove (a), suppose that x and y are both weak limits of the sequence {xn } and
set z = x − y. Then hf, zi = 0 for every f ∈ X ∗ and by Corollary 1.13, z = 0. To prove (b),
let f ∈ X ∗ and note that xn → x implies hf, xn i → hf, xi since f is continuous. To prove (c),
assume xn * x and consider the sequence {Jxn } of elements of X ∗∗ , where J : X → X ∗∗
is the bounded operator defined above. For each f ∈ X ∗ , sup |Jxn (f )| = sup |hf, xn i| < ∞
(since hf, xn i converges). By the Banach-Steinhaus Theorem, there exists a constant c such
that kxn k = kJxn k ≤ c which implies {xn } is bounded. Finally, for f ∈ X ∗
|hf, xi| = lim |hf, xn i| ≤ lim inf kf kkxn k = kf k lim inf kxn k
which implies the desired inequality since kxk = supkf k=1 |hf, xi|.
We note that in a Hilbert space H, the Riesz representation theorem implies that xn * x
means (xn , y) → (x, y) for all y ∈ H. Moreover, we have
(xn , yn ) → (x, y) if xn * x, yn → y.
This follows from the estimate
|(x, y) − (xn , yn )| = |(x − xn , y) − (xn , yn − y)| ≤ |(x − xn , y)| + kxn kky − yn k
and the fact that kxn k is bounded.
8 1. Preliminaries
Remark. In other words, every bounded equicontinuous sequence of functions has a uni-
formly convergent subsequence.
Theorem 1.21. Let X and Y be Banach spaces. If Tn : X → Y are linear and compact
for n ≥ 1 and if limn→∞ kTn − T k = 0, then T is compact. Thus, compact operators form
a closed, but not a dense, subspace of B(X, Y ).
Proof. Let {xn } be a sequence in X with M = supn kxn k < ∞. Let A1 denote an infinite
set of integers such the sequence {T1 xn }n∈A1 converges. For k ≥ 2 let Ak ⊂ Ak−1 denote
an infinite set of integers such that the sequence {Tk xn }n∈Ak converges. Choose n1 ∈ A1
and nk ∈ Ak , nk > nk−1 for k ≥ 2. Choose ε > 0. Let k be such that kT − Tk kM < ε/4
and note that
kT xni − T xnj k ≤ k(T − Tk )(xni − xnj )k + kTk xni − Tk xnj k < ε/2 + kTk xni − Tk xnj k.
Since {Tk xni }∞ ∞
i=1 converges, {T xni }i=1 is a Cauchy sequence.
Theorem 1.22. Let X and Y be normed spaces.
(a) If T ∈ B(X, Y ), then T is weakly continuous, i.e.,
xn * x implies T xn * T x.
(b) If T : X → Y is weakly continuous and X is a reflexive Banach space, then T
is bounded.
(c) If T ∈ B(X, Y ) is compact, then T is strongly continuous, i.e.,
xn * x implies T xn → T x.
1.4. Spectral Theory for Compact Linear Operators 9
Proof. Let {xn } be a sequence in H satisfying kxn k ≤ m. The sequence {T *xn } is therefore
bounded, since T * is bounded. Since T is compact, by passing to a subsequence if necessary,
we may assume that the sequence {T T *xn } converges. But then
kT *(xn − xm )k2 = (xn − xm , T T *(xn − xm ))
≤ 2mkT T *(xn − xm )k → 0 as m, n → ∞.
Since H is complete, the sequence {T *xn } is convergent and hence T * is compact.
Let T : D(T ) ⊂ H → H be a linear operator on the real Hilbert space H. The set ρ(T )
of all scalars λ ∈ R for which (T − λI)−1 ∈ B(H) is called the resolvent set of T . The
operator R(λ) = (T − λI)−1 is known as the resolvent of T . σ(T ) = R \ ρ(T ) is called the
spectrum of T . It can be shown that ρ(T ) is an open set and σ(T ) is a closed set. The
set of λ ∈ R for which there exists a nonzero x ∈ N (T − λI) is called the point spectrum
of T and is denoted by σp (T ). The elements of σp (T ) are called the eigenvalues of T and
the nonzero members of N (T − λI) are called the eigenvectors (or eigenfunctions if X
is a function space) of T .
If T is compact and λ 6= 0, then by the Fredholm alternative, either λ ∈ σp (T ) or
λ ∈ ρ(T ). Moreover, if H is infinite-dimensional, then 0 6∈ ρ(T ); otherwise, T −1 ∈ B(H)
and T −1 T = I is compact. As a consequence, σ(T ) consists of the nonzero eigenvalues of T
together with the point 0. The next result shows that σp (T ) is either finite or a countably
infinite sequence tending to zero.
Theorem 1.25. Let T : X → X be a compact linear operator on the normed space X.
Then for each r > 0 there exist at most finitely many λ ∈ σp (T ) for which |λ| > r.
1.4.3. Symmetric Compact Operators. The next result implies that a symmetric com-
pact operator on a Hilbert space has at least one eigenvalue. On the other hand, an arbi-
trary bounded, linear, symmetric operator need not have any eigenvalues. As an example,
let T : L2 (0, 1) → L2 (0, 1) be defined by T u(x) = xu(x).
Theorem 1.26. Suppose T ∈ B(H) is symmetric, i.e., (T x, y) = (x, T y) for all x, y ∈ H.
Then
kT k = sup |(T x, x)|.
kxk=1
Moreover, if H 6= {0}, then there exists a real number λ ∈ σ(T ) such that |λ| = kT k. If
λ ∈ σp (T ), then in absolute value λ is the largest eigenvalue of T .
Proof. Clearly m ≡ supkxk=1 |(T x, x)| ≤ kT k. To show kT k ≤ m, observe that for all
x, y ∈ H
2(T x, y) + 2(T y, x) = (T (x + y), x + y) − (T (x − y), x − y)
≤ m(kx + yk2 + kx − yk2 )
= 2m(kxk2 + kyk2 )
where the last step follows from the paralleogram law. Hence, if T x 6= 0 and y =
(kxk/kT xk)T x, then
2kxkkT xk = (T x, y) + (y, T x) ≤ m(kxk2 + kyk2 ) = 2mkxk2
which implies kT xk ≤ mkxk. Since this is also valid when T x = 0, we have kT k ≤ m. To
prove the ‘moreover’ part, choose xn ∈ H such that kxn k = 1 and kT k = limn→∞ |(T xn , xn )|.
By renaming a subsequence of {xn }, we may assume that (T xn , xn ) converge to some real
number λ with |λ| = kT k. Observe that
k(T − λ)xn k2 = kT xn k2 − 2λ(T xn , xn ) + λ2 kxn k2
≤ 2λ2 − 2λ(T xn , xn ) → 0.
We now claim that λ ∈ σ(T ). Otherwise, we arrive at the contradiction
1 = kxn k = k(T − λ)−1 (T − λ)xn k ≤ k(T − λ)−1 k k(T − λ)xn k → 0.
1.5. Nonlinear Functional Analysis 11
Finally, we note that if T φ = µφ, with kφk = 1, then |µ| = |(T φ, φ)| ≤ kT k which implies
the last assertion of the theorem.
and thus x is the unique fixed point. Note that the fixed point x is independent of x0 since
x is a fixed point and fixed points are unique.
Our main existence result will be based upon the following so-called method of con-
tinuity or continuation method.
12 1. Preliminaries
(1.10) x = Ts−1 (f + τ T0 x − τ T1 x) ≡ Ax
and for A : X → X we have for all x, y ∈ X
kAx − Ayk ≤ τ c(kT1 k + kT0 k)kx − yk.
By the contraction mapping theorem, (1.10) has a solution and this completes the proof.
Here h·, ·i denotes the duality pairing between Lp (Ω) and Lq (Ω).
1.5. Nonlinear Functional Analysis 13
Proof. Since u ∈ Lp (Ω), the function u(x) is measurable on Ω and thus, Pn by (i) and P
(ii), the
function f (x, u(x)) is also measurable on Ω. From the inequality ( i=1 ξi ) ≤ c ni=1 ξir
r
Remarks. (a) The following remarkable statement can be proved: If f satisfies (i) and (ii)
above and if the corresponding Nemytskii operator is such that N : Lp (Ω) → Lq (Ω), then
N is continuous, bounded and (iii) holds.
(b) If (iii) is replaced by
(iii)0 for all (x, ξ) ∈ Ω × R
|f (x, ξ)| ≤ a(x) + b|ξ|p/r
where b is a fixed nonnegative number, a ∈ Lr (Ω) is nonnegative and 1 < p, r < ∞, then
the above results are valid with q replaced by r. (i.e., N : Lp (Ω) → Lr (Ω).)
(c) We say that f satisfies the Caratheodory property, written f ∈ Car, if (i) and
(ii) above are met. If in addition (iii) is met, then we write f ∈ Car(p).
1.5.3. Differentiability. Let S be an open subset of the Banach space X. The functional
f : S ⊂ X → R is said to be Gateaux differentiable (G-diff) at a point u ∈ S if there
exists a functional g ∈ X* (often denoted by f 0 (u)) such that
d f (u + tv) − f (u)
f (u + tv) = lim = [f 0 (u)]v for all v ∈ X.
dt t=0 t→0 t
The functional f 0 (u) is called the Gateaux derivative of f at the point u ∈ S. If f is
G-diff at each point of S, the map f 0 : S ⊂ X → X* is called the Gateaux derivative of
f on S. In addition, if f 0 is continuous at u (in the operator norm), then we say that f
is C 1 at u. Note that in the case of a real-valued function of several real variables, the
Gateaux derivative is nothing more than the directional derivative of the function at u in
the direction v.
The operator B, often denoted by A0 (u), is called the Frechet derivative of A at u. Note
that if A is Frechet differentiable on S, then A0 : S → B(X, Y ). In addition, if A0 is
continuous at u (in the operator norm), we say that A is C 1 at u.
Remark. If the functional f is F-diff at u ∈ S, then it is also G-diff at u, and the two
derivatives are equal. This follows easily from the definition of the Frechet derivative. The
converse is not always true as may be easily verified by simple examples from several variable
calculus. However, if the Gateaux derivative exists in a neighborhood of u and if f ∈ C 1 at
u, then the Frechet derivative exists at u, and the two derivatives are equal.
Example 1.31. (a) Let f (ξ) ∈ C(R). Then for k ≥ 0, the corresponding Nemytskii operator
N : C k (Ω̄) → C(Ω̄) is bounded and continuous. If in addition f (ξ) ∈ C 1 (R), then N ∈ C 1
and the Frechet derivative N 0 (u) is given by
[N 0 (u)v](x) = f 0 (u(x))v(x).
Note that for u, v ∈ C k (Ω̄), |N 0 (u)v|0 ≤ |f 0 (u)|0 |v|k and so N 0 (u) ∈ B(C k (Ω̄), C ( Ω̄)) with
kN 0 (u)k ≤ |f 0 (u)|0 . Clearly N 0 (u) is continuous at each point u ∈ C k (Ω̄). Moreover,
Z 1
0 d
|N (u + v) − N u − N (u)v|0 = sup | [ f (u(x) + tv(x)) − f 0 (u(x))v(x)]dt|
x 0 dt
Z 1
≤ |v|0 sup |f 0 (u(x) + tv(x)) − f 0 (u(x))|dt.
x 0
The last integral tends to zero since f0 is uniformly continuous on compact subsets of R.
More generally, let f (ξ) ∈ C k (R). Then the corresponding Nemytskii operator N :
C k (Ω̄)
→ C k (Ω̄) is bounded and continuous. If in addition f (ξ) ∈ C k+1 (R), then N ∈ C 1
with Frechet derivative given by [N 0 (u)v](x) = f 0 (u(x))v(x). Note that |uv|k ≤ |u|k |v|k for
u, v ∈ C k (Ω̄), and since C k (Ω̄) ⊂ C(Ω̄), the Frechet derivative must be of the stated form.
(b) Let f (ξ) ∈ C k+1 (R), where k > n/2. Then we claim that the corresponding Ne-
mytskii operator N : H k (Ω) → H k (Ω) is of class C 1 with Frechet derivative given by
[N 0 (u)v](x) = f 0 (u(x))v(x).
First, suppose u ∈ C k (Ω̄). Then N (u) ∈ C k (Ω̄) by the usual chain rule. If u ∈ H k (Ω),
let um ∈ C k (Ω̄) with kum − ukk,2 → 0. Since the imbedding H k (Ω) ⊂ C(Ω̄) is continuous,
um → u uniformly, and thus f (um ) → f (u) and f 0 (um ) → f 0 (u) uniformly and hence in L2 .
Furthermore, Di f (um ) = f 0 (um )Di um → f 0 (u)Di u in L1 . Consequently, by Theorem 2.11,
we have
Di f (u) = f 0 (u)Di u.
In a similar fashion we find
Dij f (u) = f 00 (u)Di uDj u + f 0 (u)Dij u
with corresponding formulas for higher derivatives.
1.5.4. Implicit Function Theorem. The following lemmas are needed in the proof of
the implicit function theorem.
1.5. Nonlinear Functional Analysis 15
Lemma 1.32. Let S be a closed nonempty subset of the Banach space X and let M be a
metric space. Suppose A(x, λ) : S × M → S is continuous and there is a constant k < 1
such that, uniformly for all λ ∈ M
kA(x, λ) − A(y, λ)k ≤ kkx − yk for all x, y ∈ S.
Then for each λ ∈ M, A(x, λ) has a unique fixed point x(λ) ∈ S and moreover, x(λ)
depends continuously on λ.
Proof. The existence and uniqueness of the fixed point x(λ) is of course a consequence of
the contraction mapping theorem. To prove continuity, suppose λn → λ. Then
kx(λn ) − x(λ)k = kA(x(λn ), λn ) − A(x(λ), λ)k
≤ kA(x(λn ), λn ) − A(x(λ), λn )k + kA(x(λ), λn ) − A(x(λ), λ)k
≤ kkx(λn ) − x(λ)k + kA(x(λ), λn ) − A(x(λ), λ)k.
Therefore
1
kx(λn ) − x(λ)k ≤ kA(x(λ), λn ) − A(x(λ), λ)k.
1−k
By the assumed continuity of A, the right side tends to zero as n → ∞, and therefore
x(λn ) → x(λ).
Lemma 1.33. Suppose X, Y are Banach spaces. Let S ⊂ X be convex and assume A :
S → Y is Frechet differentiable at every point of S. Then
kAu − Avk ≤ ku − vk sup kA0 (w)k.
w∈S
In other words, A satisfies a Lipschitz condition with constant q = supw∈S kA0 (w)k.
Proof. For fixed u, v ∈ S, set g(t) = A(u + t(v − u)), where t ∈ [0, 1]. Using the definition
of Frechet derivative, we have
0 A(u + (t + h)(v − u)) − A(u + t(v − u))
g (t) = lim
h→0 h
0
hA (u + t(v − u))(v − u) + kh(v − u)kE
= lim
h→0 h
0
= A (u + t(v − u))(v − u).
Hence
kg(0) − g(1)k = kAu − Avk ≤ sup kg 0 (t)k
t∈[0,1]
which implies the desired result.
Lemma 1.34. Let X be a Banach space. Suppose A : B(u0 , r) ⊂ X → X is a contraction,
with Lipschitz constant q < 1, where
r ≥ (1 − q)−1 kAu0 − u0 k.
Then A has a unique fixed point u ∈ B(u0 , r).
We now consider operator equations of the form A(u, v) = 0, where A maps a subset
of X × Y into Z. For a given [u0 , v0 ] ∈ X × Y we denote the Frechet derivative of A (at
[u0 , v0 ]) with respect to the first (second) argument by Au (u0 , v0 ) (Av (u0 , v0 )).
Theorem 1.35. (Implicit Function) Let X, Y, Z be Banach spaces. For a given [u0 , v0 ] ∈
X × Y and a, b > 0, let S = {[u, v] : ku − u0 k ≤ a, kv − v0 k ≤ b}. Suppose A : S → Z
satisfies the following:
(i) A is continuous.
(ii) Av (·, ·) exists and is continuous in S (in the operator norm)
(iii) A(u0 , v0 ) = 0.
(iv) [Av (u0 , v0 )]−1 exists and belongs to B(Z, Y ).
Then there are neighborhoods U of u0 and V of v0 such that the equation A(u, v) = 0 has
exactly one solution v ∈ V for every u ∈ U . The solution v depends continuously on u.
Proof. If in S we define
B(u, v) = v − [Av (u0 , v0 )]−1 A(u, v)
it is clear that the solutions of A(u, v) = 0 and v = B(u, v) are identical. The theorem will
be proved by applying the contraction mapping theorem to B. Since
Bv (u, v) = I − [Av (u0 , v0 )]−1 Av (u, v)
Bv (·, ·) is continuous in the operator norm. Now Bv (u0 , v0 ) = 0, so for some δ > 0 there is
a q < 1 such that
kBv (u, v)k ≤ q
for ku − u0 k ≤ δ, kv − v0 k ≤ δ. By virtue of Lemma 1.33, B(u, ·) is a contraction. Since
A is continuous, B is also continuous. Therefore, since B(u0 , v0 ) = v0 , there is an ε with
0 < ε ≤ δ such that
kB(u, v0 ) − v0 k ≤ (1 − q)δ
for ku − u0 k ≤ ε. The existence of a unique fixed point in the closed ball B(v0 , δ) follows
from Lemma 1.34 and the continuity from Lemma 1.32.
Example 1.36. Let f (ξ) ∈ C 1,α (R), f (0) = f 0 (0) = 0, g(x) ∈ C α (Ω̄) and consider the
boundary value problem
(1.13) ∆u + f (u) = g(x) in Ω, u|∂Ω = 0.
Set X = Z = C α (Ω̄), Y = {u ∈ C 2,α (Ω̄) : u|∂Ω = 0} and
A(g, u) = ∆u + N (u) − g
where N is the Nemytskii operator corresponding to f . The operator A maps X × Y into
the space Z. Clearly A(0, 0) = 0 (A is C 1 by earlier examples) and
Au (0, 0)v = ∆v, v ∈ Y.
It is easily checked that all the conditions of the implicit function theorem are met. In
particular, condition (iv) is a consequence of the bounded inverse theorem. Thus, for a
function g ∈ C α (Ω̄) of sufficiently small norm (in the space C α (Ω̄)) there exists a unique
solution of (1.13) which lies near the zero function. There may, of course, be other solutions
which are not close to the zero function. (Note that the condition f 0 (0) = 0 rules out linear
functions.)
1.5. Nonlinear Functional Analysis 17
Remark. Note that the choice of X = Z = C(Ω̄), Y = {u ∈ C 2 (Ω̄) : u|∂Ω = 0} would fail
above since the corresponding linear problem is not onto. An alternate approach would be
to use Sobolev spaces. In fact, if we take X = Z = W k−2 (Ω), Y = W k (Ω) ∩ H01 (Ω) with k
sufficiently large, and if f (ξ) ∈ C k+1 (R), then as above, we can conclude the existence of a
unique solution u ∈ W k (Ω) provided kgkk−2,2 is sufficiently small. Hence, we get existence
for more general functions g; however, the solution u ∈ W k (Ω) is not a classical (i.e., C 2 )
solution in general.
1.5.5. Generalized Weierstrass Theorem. In its simplest form, the classical Weier-
strass theorem can be stated as follows: Every continuous function defined on a closed ball
in Rn is bounded and attains both its maximum and minimum on this ball. The proof
makes essential use of the fact that the closed ball is compact.
The first difficulty in trying to extend this result to an arbitrary Banach space X is
that the closed ball in X is not compact if X is infinite dimensional. However, as we shall
show, a generalized Weierstrass theorem is possible if we require a stronger property for the
functional.
A set S ⊂ X is said to be weakly closed if {un } ⊂ S, un * u implies u ∈ S, i.e.,
S contains all its weak limits. A weakly closed set is clearly closed, but not conversely.
Indeed, the set {sin nx}∞ 2
1 in L (0, π) has no limit point (because it cannot be Cauchy) so
it is closed, but zero is a weak limit that does not belong to the set. It can be shown that
every convex, closed set in a Banach space is weakly closed.
A functional f : S ⊂ X → R is weakly continuous at u0 ∈ S if for every sequence
{un } ⊂ S with un * u0 it follows that f (un ) → f (u0 ). Clearly, every functional f ∈ X*
is weakly continuous. A functional f : S ⊂ X → R is weakly lower semicontinu-
ous(w.l.s.c.) at u0 ∈ S if for every sequence {un } ⊂ S for which un * u0 it follows that
f (u0 ) ≤ lim inf n→∞ f (un ). According to Theorem 1.18, the norm on a Banach space is
w.l.s.c.. A functional f : S ⊂ X → R is weakly coercive on S if f (u) → ∞ as kuk → ∞
on S.
Theorem 1.37. Let X be a reflexive Banach space and f : C ⊂ X → R be w.l.s.c. and
assume
(i) C is a nonempty bounded weakly closed set in X or
(ii) C is a nonempty weakly closed set in X and f is weakly coercive on C.
Then
(a) inf u∈C f (u) > −∞;
(b) there is at least one u0 ∈ C such that f (u0 ) = inf u∈C f (u).
Moreover, if u0 is an interior point of C and f is G-diff at u0 , then f 0 (u0 ) = 0.
Proof. Assume (i) and let {un } ⊂ C be a minimizing sequence, i.e., limn→∞ f (un ) =
inf u∈C f (u). The existence of such a sequence follows from the definition of inf. Since X is
reflexive and C is bounded and weakly closed, there is a subsequence {un0 } and a u0 ∈ C
such that un0 * u0 . But f is w.l.s.c. and so f (u0 ) ≤ lim inf n→∞ f (un0 ) = inf u∈C f (u),
which proves (a). Since by definition, f (u0 ) ≥ inf u∈C f (u), we get (b).
Assume (ii) and fix u0 ∈ C. Since f is weakly coercive, there is a closed ball B(0, R) ⊂ X
such that u0 ∈ B ∩ C and f (u) ≥ f (u0 ) outside B ∩ C. Since B ∩ C satisfies the conditions
of (i), there is a u1 ∈ B ∩ C such that f (u) ≥ f (u1 ) for all u ∈ B ∩ C and in particular for
u0 . Thus, f (u) ≥ f (u1 ) on all of C.
18 1. Preliminaries
To prove the last statement we set ϕv (t) = f (u0 + tv). For fixed v ∈ X, ϕv (t) has a
local minimum at t = 0, and therefore hf 0 (u0 ), vi = 0 for all v ∈ X.
Remark. Even though weakly continuous functionals on closed balls attain both their inf
and sup (which follows from the above theorem), the usual functionals that we encounter
are not weakly continuous, but are w.l.s.c.. Hence this explains why we seek the inf and
not the sup in variational problems.
1.5.5.1. Convex sets. A set C in the real normed space X is called convex if (1−t)u+tv ∈ C
for all t ∈ [0, 1], u, v ∈ C. The following result is needed later (see, e.g., [?]).
Theorem 1.38. A closed convex set in a Banach space is weakly closed.
A is monotone if
hAu − Av, u − vi ≥ 0 for all u, v ∈ X.
A is strongly monotone if
hAu − Av, u − vi ≥ cku − vkpX for all u, v ∈ X
where c > 0 and p > 1.
A is coercive if
hAu, ui
lim = +∞.
kuk→∞ kuk
Remark. A strongly monotone operator is coercive. This follows immediately from hAu, ui =
hAu − A0, ui + hA0, ui ≥ ckukpX − kA0kkukX .
Proof. Set
Cr = {u ∈ C : f (u) ≤ r}.
It follows from (a) that Cr is closed and convex for all r, and thus is weakly closed (cf.
Theorem 1.38). If f is not w.l.s.c., then there is a sequence {un } ⊂ C with un * u and
f (u) > lim inf f (un ). Hence, there is an r and a subsequence {un0 } such that f (u) > r and
f (un0 ) ≤ r (i.e., un0 ∈ Cr ) for all n0 large enough. Since Cr is weakly closed, u ∈ Cr , which
is a contradiction.
Assume (b) holds and set ϕ(t) = f (u + t(v − u)). Then by Lemma 1.39, ϕ : [0, 1] → R
is convex and ϕ0 is monotone. By the classical mean value theorem,
ϕ(1) − ϕ(0) = ϕ0 (θ) ≥ ϕ0 (0), 0<θ<1
i.e.,
f (v) ≥ f (u) + hf 0 (u), v − ui for all u, v ∈ C.
If un * u, then hf 0 (u), un − ui → 0 as n → ∞. Hence, f is w.l.s.c.
In the first case we say that f is (local) maximal at u0 with respect to Mc , while in the
second case f is (local) minimal at u0 with respect to Mc . A point u0 ∈ Mc is called an
ordinary point of the manifold Mc if its F-derivative g 0 (u0 ) 6= 0.
Let u0 be an ordinary point of Mc . Then u0 is called a critical point of f with respect
to Mc if there exists a real number λ, called a Lagrange multiplier, such that
f 0 (u0 ) = λg 0 (u0 ).
As we shall see, if u0 is an extremum of f with respect to Mc , and if u0 is an ordinary point,
then u0 is a critical point of f with respect to Mc . Note that if u0 is an extremum of f with
respect to X, then we can choose λ = 0, which implies the usual result.
Lemma 1.41. Let X be a Banach space. Suppose the following hold:
(i) f, g : X → R are of class C 1
(ii) For u0 ∈ X, we can find v, w ∈ X such that
Proof. If (a) does not hold, then fix w ∈ X with g 0 (u0 )w 6= 0. By hypothesis and the above
lemma, we must have
f 0 (u0 )v · g 0 (u0 )w = f 0 (u0 )w · g 0 (u0 )v for all v ∈ X.
If we define λ = (f 0 (u0 )w)/(g 0 (u0 )w), then we obtain (b).
Remark. In applying this theorem one should be careful and not choose g(u) = kuk, since
this g is not weakly continuous.
Assume that the imbedding X ⊂ Y is compact and that the imbedding Y ⊂ Z is continuous.
Then for each ε > 0, there is a constant c(ε) such that
(1.16) kukY ≤ εkukX + c(ε)kukZ for all u ∈ X.
Proof. If for a fixed ε > 0 the inequality is false, then there exists a sequence {un } such
that
(1.17) kun kY > εkun kX + nkun kZ for all n.
Without loss of generality we can assume kun kX = 1. Since the imbedding X ⊂ Y is
compact, there is a subsequence, again denoted by {un }, with un → u in Y . This implies
un → u in Z. By (1.17), kun kY > ε and so u 6= 0. Again by (1.17), un → 0 in Z, i.e., u = 0,
which is a contradiction.
Chapter 2
Sobolev Spaces
This chapter is devoted to the study of the necessary Sobolev function spaces which permit
a modern approach to partial differential equations.
where v = Dα u. Motivated by (2.1), we now enlarge the class of functions for which the
notion of derivative can introduced.
Let u ∈ L1loc (Ω). A function v ∈ L1loc (Ω) is called the αth weak derivative of u if it
satisfies
Z Z
α |α| |α|
(2.2) uD ϕdx = (−1) vϕdx for all ϕ ∈ C0 (Ω).
Ω Ω
It can be easily shown that the weak derivative is unique. Thus we write v = Dα u to indicate
that v is the αth weak derivative of u. If a function u has an ordinary αth derivative lying
in L1loc (Ω), then it is clearly the αth weak derivative.
23
24 2. Sobolev Spaces
Example 2.1. (a) The function u(x) = |x1 | has in the ball Ω = B(0, 1) weak derivatives
ux1 = sgn x1 , uxi = 0, i = 2, . . . , n. In fact, we apply formula (2.2) as follows: For any
ϕ ∈ C01 (Ω)
Z Z Z
|x1 |ϕx1 dx = x1 ϕx1 dx − x1 ϕx1 dx
Ω Ω+ Ω−
where Ω+ = Ω ∩ (x1 > 0), Ω− = Ω ∩ (x1 < 0). Since x1 ϕ = 0 on ∂Ω and also for x1 = 0,
an application of the divergence theorem yields
Z Z Z Z
|x1 |ϕx1 dx = − ϕdx + ϕdx = − (sgn x1 )ϕdx.
Ω Ω+ Ω− Ω
|x1 |xi = 0 for i = 2, . . . , n. Note that the function |x1 | has no classical derivative with
respect to x1 in Ω.
(b) By the above computation, the function u(x) = |x| has a weak derivative u0 (x) =
sgn x on the interval Ω = (−1, 1). On the other hand, sgn x does not have a weak derivative
on Ω due to the discontinuity at x = 0.
(c) Let Ω = B(0, 1/2) ⊂ R2 and define u(x) = ln(ln(2/r)), x ∈ Ω, where r = |x| =
+ x22 )1/2 . Then u 6∈ L∞ (Ω) because of the singularity at the origin. However, we will
(x21
show that u has weak first partial derivatives.
First of all u ∈ L2 (Ω), for
Z Z 2π Z 1/2
2
|u| dx = r[ln(ln(2/r))]2 dr dθ
Ω 0 0
and a simple application of L’hopitals rule shows that the integrand is bounded and thus
the integral is finite. Similarly, it is easy to check that the classical partial derivative
− cos θ
ux1 = , where x1 = r cos θ
r ln(2/r)
also belongs to L2 (Ω). Now we show that the defining equation for the weak derivative is
met.
Let Ωε = {x : ε < r < 1/2} and choose ϕ ∈ C01 (Ω). Then by the divergence theorem
and the absolute continuity of integrals
Z Z Z Z
uϕx1 dx = lim uϕx1 dx = lim − ux1 ϕdx + uϕn1 ds
Ω ε→0 Ωε ε→0 Ωε r=ε
The same analysis applies to ux2 . Thus u has weak first partial derivatives given by the
classical derivatives which are defined on Ω\{0}.
2.1. Weak Derivatives and Sobolev Spaces 25
The space W k,p (Ω) is known as a Sobolev space of order k and power p.
We define the space W0k,p (Ω) to be the closure of the space C0k (Ω) with respect to the
norm k · kk,p . As we shall see shortly, W k,p (Ω) 6= W0k,p (Ω) for k ≥ 1. (Unless Ω = Rn .)
Remark 2.1. The spaces W k,2 (Ω) and W0k,2 (Ω) are special since they become a Hilbert
space under the inner product
Z X
(u, v)k,2 = (u, v)W k,2 (Ω) = Dα uDα vdx.
Ω |α|≤k
Since we shall be dealing mostly with these spaces in the sequel, we introduce the special
notation:
H k (Ω) = W k,2 (Ω), H0k (Ω) = W0k,2 (Ω).
Theorem 2.2. W k,p (Ω) is a Banach space under the norm (2.3). If 1 < p < ∞, it is
reflexive, and if 1 ≤ p < ∞, it is separable.
Proof. 1. We first prove that W k,p (Ω) is complete with respect to the norm (2.3). We
prove this for 1 ≤ p < ∞; the case p = ∞ is similar. Let {un } be a Cauchy sequence of
elements in W k,p (Ω), i.e.,
p
X Z
kun − um kk,p = |Dα un − Dα um |p dx → 0 as m, n → ∞.
|α|≤k Ω
to Lp (Ω), a simple limit argument shows that uα is the αth weak derivative of u0 . In fact,
Z Z Z Z
α α |α| α |α|
uD ϕdx ← un D ϕdx = (−1) ϕD un dx → (−1) uα ϕdx.
Ω Ω Ω Ω
Hence u0
∈ W k,p (Ω)
and kun − u0 k
k,p → 0 as n → ∞. This proves the completeness of
k,p
W (Ω); hence it is a Banach space.
2. Consider the map T : W 1,p (Ω) → (Lp (Ω))n+1 defined by
T u = (u, D1 u, . . . , Dn u).
26 2. Sobolev Spaces
for v = (v1 , . . . , vn+1 ) ∈ (Lp (Ω))n+1 , then T is a (linear) isometry. Now (Lp (Ω))n+1 is
reflexive for 1 < p < ∞ and separable for 1 ≤ p < ∞. Since W 1,p (Ω) is complete, its image
under the isometry T is a closed subspace of (Lp (Ω))n+1 which inherits the corresponding
properties as does W 1,p (Ω) (see Theorem 1.14). Similarly, we can handle the case k ≥ 2.
the latter equality being valid since ωh vanishes outside the (open) ball B(x, h). Thus
the values of uh (x) depend only on the values of u on the ball B(x, h). In particular, if
dist(x, supp(u)) ≥ h, then uh (x) = 0.
Theorem 2.5. Let Ω be a nonempty open set in Rn . Then
(a) uh ∈ C ∞ (Rn ).
(b) If supp(u) is a compact subset of Ω, then uh ∈ C0∞ (Ω) for all h sufficiently
small.
With respect to a bounded set Ω we construct another set Ω(h) as follows: with each
point x ∈ Ω as center, draw a ball of radius h; the union of these balls is then Ω(h) . Clearly
Ω(h) ⊃ Ω. Moreover, uh can be different from zero only in Ω(h) .
28 2. Sobolev Spaces
Corollary 2.6. Let Ω be a nonempty bounded open set in Rn and let h > 0 be any number.
Then there exists a function η ∈ C ∞ (Rn ) such that
0 ≤ η(x) ≤ 1; η(x) = 1, x ∈ Ω(h) ; η(x) = 0, x ∈ (Ω(3h) )c .
Such a function is called a cut-off function for Ω.
Proof. Let χ(x) be the characteristic function of the set Ω(2h) : χ(x) = 1 for x ∈ Ω(2h) , χ(x) =
0 for x 6∈ Ω(2h) and set
Z
η(x) ≡ χh (x) = ωh (x − y)χ(y)dy.
Rn
Then Z
η(x) = ωh (x − y)dy ∈ C ∞ (Rn ),
Ω(2h)
Z
0 ≤ η(x) ≤ ωh (x − y)dy = 1,
Rn
and
(R
B(x,h) ωh (x − y)dy = 1, x ∈ Ω(h) ,
Z
η(x) = ωh (x − y)χ(y)dy =
B(x,h) 0, x ∈ (Ω(3h) )c .
In particular, we note that if Ω0 ⊂⊂ Ω, there is a function η ∈ C0∞ (Ω) such that η(x) = 1
for x ∈ Ω0 , and 0 ≤ η(x) ≤ 1 in Ω.
Henceforth, the notation Ω0 ⊂⊂ Ω means that Ω0 , Ω are open sets, Ω0 is bounded, and
that Ω0 ⊂ Ω.
We need the following well-known result.
Theorem 2.7. (Partition of Unity) Assume Ω ⊂ Rn is bounded and Ω ⊂⊂ ∪N i=1 Ωi ,
where each Ωi is open. Then there exist C ∞ functions ψi (x) (i = 1, . . . , N ) such that
(a) 0 ≤ ψi (x) ≤ 1
(b) ψi has its support in Ωi
PN
(c) i=1 ψi (x) = 1 for every x ∈ Ω.
Proof. Choose a > 0 large enough so that Ω is strictly contained in the ball B(0, a). Then
the function
u(x) if x ∈ Ω,
U (x) =
0 if x ∈ B(0, 2a) \ Ω
belongs to Lp (B(0, 2a)). For ε > 0, there is a function Ū ∈ C(B̄(0, 2a)) which satisfies the
inequality kU − Ū kLp (B(0,2a)) < ε/3. By multiplying Ū by an appropriate cut-off function,
it can be assumed that Ū (x) = 0 for x ∈ B(0, 2a) \ B(0, a). Therefore for |z| ≤ a,
kU (x + z) − Ū (x + z)kLp (B(0,2a)) = kU (x) − Ū (x)kLp (B(0,a)) ≤ ε/3.
2.2. Approximations and Extensions 29
Since function Ū is uniformly continuous in B(0, 2a), there is a 0 < δ < a such that
kŪ (x + z) − Ū (x)kLp (B(0,2a)) ≤ ε/3 whenever |z| < δ. Hence for |z| < δ we easily see that
ku(x + z) − u(x)kLp (Ω) = kU (x + z) − U (x)kLp (B(0,2a)) ≤ ε.
Theorem 2.9. Let Ω be a nonempty open set in Rn . If u ∈ Lp (Ω) (1 ≤ p < ∞), then
(a) kuh kp ≤ kukp
(b) kuh − ukp → 0 as h → 0.
If u ∈ C k (Ω̄) and Ω̄ is compact, then, for all Ω0 ⊂⊂ Ω,
(c) kuh − ukC k (Ω̄0 ) → 0 as h → 0.
1/p 1/q
Proof. 1. If 1 < p < ∞, let q = p/(p − 1). Then ωh = ωh ωh and Hölder’s inequality
implies
Z Z p/q
p p
|uh (x)| ≤ ωh (x − y)|u(y)| dy ωh (x − y)dy
Ω Ω
Z
≤ ωh (x − y)|u(y)|p dy
Ω
which obviously holds also for p = 1. An application of Fubini’s Theorem gives
Z Z Z Z
p p
|uh (x)| dx ≤ ωh (x − y)dx |u(y)| dy ≤ |u(y)|p dy
Ω Ω Ω Ω
The right-hand side goes to zero as h → 0 since every u ∈ Lp (Ω) is p-mean continuous.
3. We now prove (c) for k = 0. Let Ω0 , Ω00 be such that Ω0 ⊂⊂ Ω00 ⊂⊂ Ω. Let h0 be the
shortest distance between ∂Ω0 and ∂Ω00 . Take h < h0 . Then
Z
uh (x) − u(x) = [u(y) − u(x)]ωh (x − y)dy.
B(x,h)
30 2. Sobolev Spaces
If x ∈ Ω̄0 , then in the above integral y ∈ Ω̄00 . Now u is uniformly continuous in Ω̄00 and
ωh ≥ 0, and therefore for an arbitrary ε > 0 we have
Z
|uh (x) − u(x)| ≤ ε ωh (x − y)dy = ε
B(x,h)
Proof. Suppose first that Ω is bounded and let Ω0 ⊂⊂ Ω. For a given u ∈ Lp (Ω) set
u(x), x ∈ Ω0
v(x) =
0, x ∈ Ω\Ω0 .
Then Z Z
|u − v|p dx = |u|p dx.
Ω Ω\ Ω0
By the absolute continuity of integrals, we can choose Ω0 so that the integral on the right
is arbitrarily small, i.e., ku − vkp < ε/2. Since supp(v) is a compact subset of Ω, Theorems
2.5(b) and 2.9(b) imply that for h sufficiently small, vh (x) ∈ C0∞ (Ω) with kv − vh kp < ε/2,
and therefore ku − vh kp < ε. If Ω is unbounded, choose a ball B large enough so that
Z
|u|p dx < ε/2
Ω\Ω0
0
where Ω = Ω ∩ B, and repeat the proof just given.
= ωh (x − y) · v(y)dy = vh (x).
Ω00
2.2. Approximations and Extensions 31
We now note some properties of W k,p (Ω) which follow easily from the results of this
and the previous section.
(a) If Ω0 ⊂ Ω and if u ∈ W k,p (Ω), then u ∈ W k,p (Ω0 ).
(b) If u ∈ W k,p (Ω) and |a(x)|k,∞ < ∞, then au ∈ W k,p (Ω). In this case any weak
derivative Dα (au) is computed according to the usual rule of differentiating the
product of functions.
(c) If u ∈ W k,p (Ω) and uh is its mollified function, then for any compact set S ⊂
Ω, kuh − ukW k,p (S) → 0 as h → 0. If in addition, u has compact support in Ω, then
kuh − ukk,p → 0 as h → 0.
More generally, we have the following global approximation theorems. (See Meyers and
Serrin H = W . The proofs make use of a partition of unity argument.)
Theorem 2.13. Assume Ω is bounded and let u ∈ W k,p (Ω), 1 ≤ p < ∞. Then there exist
functions um ∈ C ∞ (Ω) ∩ W k,p (Ω) such that
um → u in W k,p (Ω).
In other words, C ∞ (Ω) ∩ W k,p (Ω) is dense in W k,p (Ω).
Theorem 2.14. Assume Ω is bounded and ∂Ω ∈ C 1 . Let u ∈ W k,p (Ω), 1 ≤ p < ∞. Then
there exist functions um ∈ C ∞ (Ω̄) such that
um → u in W k,p (Ω).
In other words, C ∞ (Ω̄) is dense in W k,p (Ω).
Exercise 2.6. Prove the product rule for weak derivatives:
Di (uv) = (Di u)v + u(Di v)
where u, Di u are locally Lp (Ω), v, Di v are locally Lq (Ω) (p > 1, 1/p + 1/q = 1).
Exercise 2.7. (a) If u ∈ W0k,p (Ω) and v ∈ C k (Ω̄), prove that uv ∈ W0k,p (Ω).
(b) If u ∈ W k,p (Ω) and v ∈ C0k (Ω), prove that uv ∈ W0k,p (Ω).
32 2. Sobolev Spaces
Proof. 1. According to Theorem 2.11, there exists a sequence {uh } ⊂ C 1 (Ω) such that
kuh − ukL1 (Ω0 ) → 0, kDα uh − Dα ukL1 (Ω0 ) → 0 as h → 0, where Ω0 ⊂⊂ Ω. Thus
Z Z
0
|f (uh ) − f (u)|dx ≤ sup |f | |uh − u|dx → 0 as h → 0
Ω0 Ω0
Z Z
0 α 0 α 0
|f (uh )D uh − f (u)D u|dx ≤ sup |f | |Dα uh − Dα u|dx
Ω0 Ω0
Z
+ |f 0 (uh ) − f 0 (u)||Dα u|dx.
Ω0
Since kuh − ukL1 (Ω0 ) → 0, there exists a subsequence of {uh }, which we call {uh } again,
which converges a.e. in Ω0 to u. Moreover, since f 0 is continuous, {f 0 (uh )} converges to f 0 (u)
a.e. in Ω0 . Hence the last integral tends to zero by the dominated convergence theorem.
Consequently, the sequences {f (uh )}, {f 0 (uh )Dα uh } tend to f (u), f 0 (u)Dα u respectively,
and the first conclusion follows by an application of Theorem 2.11 again.
2. If f (0) = 0, the mean value theorem implies |f (s)| ≤ M |s| for all s ∈ R. Thus,
|f (u(x))| ≤ M |u(x)| for all x ∈ Ω and so f ◦ u ∈ Lp (Ω) if u ∈ Lp (Ω). Similarly,
f 0 (u(x))Dα u ∈ Lp (Ω) if u ∈ W 1,p (Ω), which shows that f ◦ u ∈ W 1,p (Ω).
Corollary 2.16. Let Ω be a bounded open set in Rn . If u has an αth weak derivative
Dα u, |α| = 1, then so does |u| and
Dα u
if u > 0
α
D |u| = 0 if u = 0
−Dα u if u < 0
i.e., Dα |u| = (sgn u)Dα u for u 6= 0. In particular, if u ∈ W 1,p (Ω), then |u| ∈ W 1,p (Ω).
where
Dα u if u > 0
v=
0 if u ≤ 0
Proof. Given u ∈ W01,p (Ω), let un ∈ C01 (Ω) with kun − uk1,p → 0 and define vn = f ◦ un .
Since un has compact support and f (0) = 0, vn has compact support. Also vn is Lipschitz
continuous, for
Hence vn ∈ Lp (Ω). Since vn is absolutely continuous on any line segment in Ω, its par-
tial derivatives (which exist almost everywhere) coincide almost everywhere with the weak
derivatives. Moreover, we see from above that |∂vn /∂xi | ≤ cn for 1 ≤ i ≤ n, and as Ω is
bounded, ∂vn /∂xi ∈ Lp (Ω). Thus vn ∈ W 1,p (Ω) and has compact support, which implies
vn ∈ W01,p (Ω). From the relation
it follows that kvn − f ◦ ukp → 0. Furthermore, if ei is the standard ith basis vector in Rn ,
we have
|vn (x + hei ) − vn (x)| |un (x + hei ) − un (x)|
≤c
|h| |h|
and so
∂vn ∂un
lim sup k kp ≤ c lim sup k kp .
n→∞ ∂xi n→∞ ∂xi
But, {∂un /∂xi } is a convergent sequence in Lp (Ω) and therefore {∂vn /∂xi } is bounded in
Lp (Ω) for each 1 ≤ i ≤ n. Since kvn k1,p is bounded and W01,p (Ω) is reflexive, a subsequence
of {vn } converges weakly in W 1,p (Ω), and thus weakly in Lp (Ω) to some element of W01,p (Ω).
Thus, f ◦ u ∈ W01,p (Ω).
Proof. We apply the preceding theorem with f (t) = |t|. Thus |u| ∈ W01,p (Ω). Now
u+ = (|u| + u)/2 and u− = (u − |u|)/2. Thus u+ , u− ∈ W01,p (Ω).
34 2. Sobolev Spaces
2.2.4. Extensions. If Ω ⊂ Ω0 , then any function u(x) ∈ C0k (Ω) has an obvious extension
U (x) ∈ C0k (Ω0 ). From the definition of W0k,p (Ω) it follows that the function u(x) ∈ W0k,p (Ω)
and extended as being equal to zero in Ω0 \Ω belongs to W0k,p (Ω0 ). In general, a function
u ∈ W k,p (Ω) and extended by zero to Ω0 will not belong to W k,p (Ω0 ). (Consider the function
u(x) ≡ 1 in Ω.) However, if u ∈ W k,p (Ω) has compact support in Ω, then u ∈ W0k,p (Ω) and
thus the obvious extension belongs to W0k,p (Ω0 ).
Proof. 1. Suppose first that u ∈ C k (Ω̄). Let y = ψ(x) define a C k diffeomorphism that
straightens the boundary near x0 = (x01 , . . . , x0n ) ∈ ∂Ω. In particular, we assume there is a
ball B = B(x0 , r) such that ψ(B ∩ Ω) ⊂ Rn+ (i.e., yn > 0), ψ(B ∩ ∂Ω) ⊂ ∂Rn+ . (e.g., we
could choose yi = xi − x0i for i = 1, . . . , n − 1 and yn = xn − ϕ(x1 , . . . , xn−1 ), where ϕ is of
class C k . Moreover, without loss of generality, we can assume yn > 0 if x ∈ B ∩ Ω.)
2. Let G and G+ = G ∩ Rn+ be respectively, a ball and half-ball in the image of ψ such
that ψ(x0 ) ∈ G. Setting ū(y) = u ◦ ψ −1 (y) and y = (y1 , . . . , yn−1 , yn ) = (y 0 , yn ), we define
an extension Ū (y) of ū(y) into yn < 0 by
k+1
X
Ū (y 0 , yn ) = ci ū(y 0 , −yn /i), yn < 0
i=1
where the ci are constants determined by the system of equations
k+1
X
(2.5) ci (−1/i)m = 1, m = 0, 1, . . . , k.
i=1
Note that the determinant of the system (2.5) is nonzero since it is the Vandemonde deter-
minant. One verifies readily that the extended function Ū is continuous with all derivatives
up to order k in G. For example,
k+1
X
lim Ū (y) = ci ū(y 0 , 0) = ū(y 0 , 0)
y→(y 0 ,0)
i=1
2.2. Approximations and Extensions 35
Finally
k+1
X
lim Ūyn (y) = ci (−1/i)ūyn (y 0 , 0) = ūyn (y 0 , 0)
y→(y 0 ,0)
i=1
by virtue of (2.5) with m = 1. Similarly we can handle the higher derivatives. Thus
w = Ū ◦ ψ ∈ C k (B 0 ) for some ball B 0 = B 0 (x0 ) and w = u in B 0 ∩ Ω, (If x ∈ B 0 ∩ Ω, then
ψ(x) ∈ G+ and w(x) = Ū (ψ(x)) = ū(ψ(x)) = u(ψ −1 ψ(x)) = u(x)) so that w provides a C k
extension of u into Ω ∪ B 0 . Moreover,
sup |ū(y)| = sup |u(ψ −1 (y))| ≤ sup |u(x)|
G+ G+ Ω
Since a similar computation for the derivatives holds, it follows that there is a constant
c > 0, independent of u, such that
kwkC k (Ω̄∪B 0 ) ≤ ckukC k (Ω̄) .
which implies that {Um } is a Cauchy sequence and so converges to a U ∈ W0k,p (Ω00 ), since
Um ∈ C0k (Ω00 ). Now extend Um , U by 0 to Ω0 . It is easy to see that U is the desired
extension.
5. We now prove (c). At any point x0 ∈ ∂Ω let the mapping ψ and the ball G be defined
as in (b). By definition, u ∈ C k (∂Ω) implies that ū = u ◦ ψ −1 ∈ C k (G ∩ ∂Rn+ ). We define
Φ̄(y 0 , yn ) = ū(y 0 ) in G and set Φ(x) = Φ̄ ◦ ψ(x) for x ∈ ψ −1 (G). Clearly, Φ ∈ C k (B̄) for
some ball B = B(x0 ) and Φ = u on B ∩ ∂Ω. Now let {Bi } be a finite covering of ∂Ω by
balls such as B and let Φi be the corresponding C k functions defined on Bi . For each i, we
define the function Ui (x) as follows: in the ball Bi take it equal to Φi , outside Bi take it
equal to zero if x 6∈ ∂Ω and equal to u(x) if x ∈ ∂Ω. The proof can now be completed as in
(b) by use of an appropriate partition of unity.
36 2. Sobolev Spaces
2.2.5. Trace Theorem. Unless otherwise stated, Ω will denote a bounded open connected
set in Rn , i.e., a bounded domain. Let Γ be a surface which lies in Ω̄ and has the represen-
tation
xn = ϕ(x0 ), x0 = (x1 , . . . , xn−1 )
where ϕ(x0 ) is Lipschitz continuous in Ū . Here U is the projection of Γ onto the coordinate
plane xn = 0. Let p ≥ 1. A function u defined on Γ is said to belong to Lp (Γ) if
Z
1
kukLp (Γ) ≡ ( |u(x)|p dS) p < ∞
Γ
where
Z Z n−1
p 0 0 p
X ∂ϕ 0 2 1 0
|u(x)| dS = |u(x , ϕ(x ))| [1 + ( (x )) ] 2 dx .
Γ U ∂xi
i=1
Thus Lp (Γ) reduces to a space of the type Lp (U ) where U is a domain in Rn−1 .
For every function u ∈ C(Ω̄), its values γ0 u ≡ u|Γ on Γ are uniquely given. The function
γ0 u will be called the trace of the function u on Γ. Note that u ∈ Lp (Γ) since γ0 u ∈ C(Γ).
On the other hand, if we consider a function u defined a.e. in Ω (i.e., functions are
considered equal if they coincide a.e.), then the values of u on Γ are not uniquely determined
since meas(Γ) = 0. In particular, since ∂Ω has measure 0, there exist infinitely many
extensions of u to Ω̄ that are equal a.e. We shall therefore introduce the concept of trace for
functions in W 1,p (Ω) so that if in addition, u ∈ C(Ω̄), the new definition of trace reduces
to the definition given above.
Lemma 2.20. Let ∂Ω ∈ C 0,1 . Then for u ∈ C 1 (Ω̄),
(2.6) kγ0 ukLp (∂Ω) ≤ ckuk1,p
where the constant c > 0 does not depend on u.
Proof. For simplicity, let n = 2. The more general case is handled similarly. In a neigh-
borhood of a boundary point x ∈ ∂Ω, we choose a local (ξ, η)-coordinate system, where the
boundary has the local representation
η = ϕ(ξ), −α ≤ ξ ≤ α
with the C 0,1 function ϕ. Then there exists a β > 0 such that all the points (ξ, η) with
−α ≤ ξ ≤ α, ϕ(ξ) − β ≤ η ≤ ϕ(ξ)
belong to Ω̄. Let u ∈ C 1 (Ω̄). Then
Z ϕ(ξ)
u(ξ, ϕ(ξ)) = uη (ξ, η)dη + u(ξ, t)
t
where ϕ(ξ) − β ≤ t ≤ ϕ(ξ). Applying the inequality (a + b)p ≤ 2p−1 (ap + bp ) together with
Hölder’s inequality we have
Z ϕ(ξ)
|u(ξ, ϕ(ξ))|p ≤ 2p−1 β p−1 |uη (ξ, η)|p dη + 2p−1 |u(ξ, t)|p .
ϕ(ξ)−β
where S denotes a local boundary strip. Suppose ϕ(·) is C 1 . Then the differential of arc
length is given by ds = (1 + ϕ02 )1/2 dξ. Addition of the local inequalities (2.7) yields the
assertion (2.6). Now if ϕ(·) is merely Lipschitz continuous, then the derivative ϕ0 exists a.e.
and is bounded. Thus we also obtain (2.6).
Since C 1 (Ω̄) = W 1,p (Ω), the bounded linear operator γ0 : C 1 (Ω̄) ⊂ W 1,p (Ω) → Lp (∂Ω)
can be uniquely extended to a bounded linear operator γ0 : W 1,p (Ω) → Lp (∂Ω) such that
(2.6) remains true for all u ∈ W 1,p (Ω). More precisely, we obtain γ0 u in the following
way: Let u ∈ W 1,p (Ω). We choose a sequence {un } ⊂ C 1 (Ω̄) with kun − uk1,p → 0. Then
kγ0 un − γ0 ukLp (∂Ω) → 0.
The function γ0 u (as an element of Lp (∂Ω)) will be called the trace of the function
u ∈ W 1,p (Ω) on the boundary ∂Ω. (kγ0 ukLp (∂Ω) will be denoted by kukLp (∂Ω) .) Thus the
trace of a function is defined for any element u ∈ W 1,p (Ω).
Proof. 1. Suppose u ∈ C(Ω̄) ∩ W 1,p (Ω). Then by Theorem 2.19, u can be extended into
Ω0 (Ω ⊂⊂ Ω0 ) such that its extension U ∈ C(Ω̄0 ) ∩ W 1,p (Ω0 ). Let Uh (x) be the mollified
function for U . Since Uh → U as h → 0 in both the norms k · kC(Ω̄) , k · kW 1,p (Ω) , we find that
as h → 0, Uh |∂Ω → u|∂Ω uniformly and Uh |∂Ω → γ0 u in Lp (∂Ω). Consequently, γ0 u = u|∂Ω .
2. Now au ∈ W 1,p (Ω) if a ∈ C 1 (Ω̄), u ∈ W 1,p (Ω) and consequently, γ0 (au) is defined.
Let {un } ⊂ C 1 (Ω̄) with kun − uk1,p → 0. Then
γ0 (aun ) = γ0 a · γ0 un
and the desired product formula follows by virtue of the continuity of γ0 .
3. If u ∈ W01,p (Ω), then there is a sequence {un } ⊂ C01 (Ω) with kun − uk1,p → 0.
But un |∂Ω = 0 and as n → ∞, un |∂Ω → γ0 u in Lp (∂Ω) which implies γ0 u = 0. Hence
W01,p (Ω) ⊂ N (γ0 ). Now suppose u ∈ N (γ0 ). If u ∈ W 1,p (Ω) has compact support in Ω,
then by an earlier remark, u ∈ W01,p (Ω). If u does not have compact support in Ω, then it
can be shown that there exists a sequence of cut-off functions ηk such that ηk u ∈ W 1,p (Ω)
has compact support in Ω, and moreover, kηk u − uk1,p → 0. By using the corresponding
mollified functions, it follows that u ∈ W01,p (Ω) and N (γ0 ) ⊂ W01,p (Ω). Details can be found
in Evans’s book.
4. To see that R(γ0 ) = Lp (∂Ω), let f ∈ Lp (∂Ω) and let ε > 0 be given. Then there is a
u ∈ C 1 (∂Ω) such that ku − f kLp (∂Ω) < ε. If we let U ∈ C 1 (Ω̄) be the extension of u into Ω̄,
then clearly kγ0 U − f kLp (∂Ω) < ε, which is the desired result since U ∈ W 1,p (Ω).
Remark 2.8. We note that the function u ≡ 1 belongs to W 1,p (Ω) ∩ C(Ω̄) and its trace
on ∂Ω is 1. Hence this function does not belong to W01,p (Ω), which establishes the earlier
assertion that W01,p (Ω) 6= W 1,p (Ω).
38 2. Sobolev Spaces
Let u ∈ W k,p (Ω), k > 1. Since any weak derivative Dα u of order |α| < k belongs to
W 1,p (Ω), this derivative has a trace γ0 Dα u belonging to Lp (∂Ω). Moreover
kDα ukLp (∂Ω) ≤ ckDα uk1,p ≤ ckukk,p
for constant c > 0 independent of u.
Assuming the boundary ∂Ω ∈ C 1 , the unit outward normal vector n to ∂Ω exists and
is bounded. Thus, the concept of traces makes it possible to introduce, for k ≥ 2, ∂u/∂n
for u ∈ W k,p (Ω). More precisely, for k ≥ 2, there exist traces of the functions u, Di u so
that, if ni are the direction cosines of the normal, we may define
Xn
γ1 u = (γ0 (Di u))ni , u ∈ W k,p (Ω), k ≥ 2.
i=1
The trace operator γ1 : W k,p (Ω) → Lp (∂Ω) is continuous and γ1 u = (∂u/∂n)|∂Ω for u ∈
C 1 (Ω̄) ∩ W k,p (Ω).
For a function u ∈ C k (Ω̄) we define the various traces of normal derivatives given by
∂j u
|∂Ω , 0 ≤ j ≤ k − 1.
γj u =
∂nj
Each γj can be extended by continuity to all of W k,p (Ω) and we obtain the following:
Theorem 2.22. (Higher-order traces) Suppose ∂Ω ∈ C k . Then there is a unique con-
Qk−1 k−1−j,p
tinuous linear operator γ = (γ0 , γ1 , . . . , γk−1 ) : W k,p (Ω) → j=0 W (∂Ω) such that
k
for u ∈ C (Ω̄)
∂j u
γ0 u = u|∂Ω , γj u = |∂Ω , j = 1, . . . , k − 1.
∂nj
Moreover, N (γ) = W0k,p (Ω) and R(γ) = k−1 k−1−j,p (∂Ω).
Q
j=0 W
The Sobolev spaces W k−1−j,p (∂Ω), which are defined over ∂Ω, can be defined locally.
Proof. Let {un } and {vn } be sequences of functions in C 1 (Ω̄) with kun − ukH 1 (Ω) →
0, kvn − vkH 1 (Ω) → 0 as n → ∞. Formula (2.8) holds for un , vn
Z Z Z
vn Di un dx = un vn ni dS − un Di vn dx
Ω ∂Ω Ω
and upon letting n → ∞ relation (2.8) follows.
Corollary 2.24. Let ∂Ω ∈ C 1 .
(a) If v ∈ H 1 (Ω) and u ∈ H 2 (Ω) then
Z Z Z
v∆udx = γ0 v · γ1 udS − (∇u · ∇v)dx (Green’s 1st identity).
Ω ∂Ω Ω
2.3. Sobolev embedding Theorems 39
Proof. If in (2.8) we replace u by Di u and sum from 1 to n, then Green’s 1st identity is
obtained. Interchanging the roles of u, v in Green’s 1st identity and subtracting the two
identities yields Green’s 2nd identity.
Exercise 2.9. Establish the following one-dimensional version of the trace theorem: If
u ∈ W 1,p (Ω), where Ω = (a, b), then
kukLp (∂Ω) ≡ (|u(a)|p + |u(b)|p )1/p ≤ const kukW 1,p (Ω)
where the constant is independent of u.
The above results are valid for W0k,p (Ω) spaces on arbitrary bounded domains Ω.
A series of special results will be needed to prove the above theorem. Only selected
proofs will be given to illustrate some of the important techniques.
Proof. First assume p = 1. Since u has compact support, for each i = 1, . . . , n we have
Z xi
u(x) = uxi (x1 , . . . , xi−1 , yi , xi+1 , . . . , xn )dyi
−∞
and so Z ∞
|u(x)| ≤ |∇u(x1 , . . . , yi , . . . , xn )|dyi (i = 1, . . . , n).
−∞
2.3. Sobolev embedding Theorems 41
Consequently
n Z 1
n Y ∞
n−1
(2.17) |u(x)| n−1 ≤ |∇u(x1 , . . . , yi , . . . , xn )|dyi .
i=1 −∞
the last inequality resulting from the extended Hölder inequality in the appendix.
We continue by integrating with respect to x2 , . . . , xn and applying the extended Hölder
inequality to eventually find (pull out an integral at each step)
Z n Z ∞ Z ∞ 1
n Y n−1
|u(x)| n−1 dx ≤ ··· |∇u|dx1 . . . dyi . . . dxn
Rn i=1 −∞ −∞
Z n
n−1
= |∇u|dx
Rn
which is estimate (2.16) for p = 1.
Consider now the case that 1 < p < n. We shall apply the last estimate to v = |u|γ ,
where γ > 1 is to be selected. First note that
(γuγ−1 Di u)2
if u ≥ 0
(Di |u|γ )2 = = (γ|u|γ−1 Di u)2 .
(−γ(−u)γ−1 Di u)2 if u ≤ 0
Thus v ∈ C01 (Rn ), and
Z n−1 Z
γn n
|u(x)| n−1 dx ≤ |∇|u|γ |dx
Rn Rn
Z
= γ |u|γ−1 |∇u|dx
Rn
Z p−1 Z 1
p(γ−1) p p
p
≤ γ |u| p−1 dx |∇u| dx .
Rn Rn
We set
p(n − 1)
γ= >1
n−p
in which case
γn p(γ − 1) np
= = = p∗ .
n−1 p−1 n−p
Thus, the above estimate becomes
Z 1∗ Z 1
p p
p∗ p
|u| dx ≤C |∇u| dx .
Rn Rn
42 2. Sobolev Spaces
Theorem 2.27. Let Ω ⊂ Rn be bounded and open, with ∂Ω ∈ C 1 . Assume 1 ≤ p < n, and
∗
u ∈ W 1,p (Ω). Then u ∈ Lp (Ω) and
(2.18) kukLp∗ (Ω) ≤ CkukW 1,p (Ω)
where the constant C depends only on p,n and Ω.
Theorem 2.28. Let Ω ⊂ Rn be bounded and open. Assume 1 ≤ p < n, and u ∈ W01,p (Ω).
Then u ∈ Lq (Ω) and
kukLq (Ω) ≤ Ck∇ukLp (Ω)
for each q ∈ [1, p∗ ], the constant C depending only on p,q,n and Ω.
Proof. Since u ∈ W01,p (Ω), there are functions um ∈ C0∞ (Ω) such that um → u in W 1,p (Ω).
We extend each function um to be 0 in Rn \Ω̄ and apply Lemma 2.26 to discover (as above)
kukLp∗ (Ω) ≤ Ck∇ukLp (Ω) .
Since |Ω| < ∞, we furthermore have
kukLq (Ω) ≤ CkukLp∗ (Ω)
for every q ∈ [1, p∗ ].
2.3.2. Morrey’s Inequality. We now turn to the case n < p < ∞. The next result shows
that if u ∈ W 1,p (Ω), then u is in fact Hölder continuous, after possibly being redefined on
a set of measure zero.
Theorem 2.29. (Morrey’s Inequality) Assume n < p < ∞. Then there exists a constant
C, depending only on p and n, such that
(2.20) kuk 0,1− n ≤ CkukW 1,p (Rn ) , ∀ u ∈ C 1 (Rn ).
C p (Rn )
Proof. We first prove the following inequality: for all x ∈ Rn , r > 0 and all u ∈ C 1 (Rn ),
rn |∇u(y)|
Z Z
(2.21) |u(y) − u(x)| dy ≤ dy.
B(x,r) n B(x,r) |x − y|n−1
2.3. Sobolev embedding Theorems 43
To prove this, note that, for any w with |w| = 1 and 0 < s < r,
Z s
d
|u(x + sw) − u(x)| = u(x + tw)dt
0 dt
Z s
= ∇u(x + tw) · wdt
0
Z s
≤ |∇u(x + sw)| dt.
0
Now we integrate w over ∂B(0, 1) to obtain
Z Z sZ
|u(x + sw) − u(x)| dS ≤ |∇u(x + sw)| dSdt
∂B(0,1) 0 ∂B(0,1)
|∇u(y)|
Z
= dy
B(x,s) |x − y|n−1
|∇u(y)|
Z
≤ dy.
B(x,r) |x − y|n−1
Multiply both sides by sn−1 and integrate over s ∈ (0, r) and we obtain (2.21).
To establish the bound on kukC 0 (Rn ) , we observe that, by (2.21), for x ∈ Rn ,
Z Z
1 1
|u(x)| ≤ |u(y) − u(x)| dy + |u(y)| dy
|B(x, 1)| B(x,1) |B(x, 1)| B(x,1)
! p−1
Z Z1/p (1−n)p
p
This inequality and the bound on kukC 0 above complete the proof.
Theorem 2.30. (Estimates for W 1,p , n < p ≤ ∞) Let Ω ⊂ Rn be bounded and open,
with ∂Ω ∈ C 1 . Assume n < p < ∞, and u ∈ W 1,p (Ω). Then, after possibly redefining u on
0,1− n
a null set, u ∈ C p (Ω̄) and
kuk 0,1− n
p
≤ CkukW 1,p (Ω)
C (Ω̄)
Moreover, since U has compact support, there exist mollified functions um ∈ C0∞ (Rn ) such
that um → U in W 1,p (Rn ) (and hence on compact subsets). Now according to Morrey’s
inequality,
kum − ul kC 0,1−n/p (Rn ) ≤ Ckum − ul kW 1,p (Rn )
2.3.3. General Cases. We can now concatenate the above estimates to obtain more com-
plicated inequalities.
Assume kp < n and u ∈ W k,p (Ω). Since Dα u ∈ Lp (Ω) for all |α| ≤ k, the Sobolev-
Nirenberg-Gagliardo inequality implies
2.4. Compactness
We now consider the compactness of the embeddings. Note that if X and Y are Banach
spaces with X ⊂ Y then we say that X is compactly embedded in Y , written X ⊂⊂ Y ,
provided
(i) kukY ≤ CkukX (u ∈ X) for some constant C; that is, the embedding is continuous;
(ii) each bounded sequence in X has a convergent subsequence in Y .
Before we present the next result we recall some facts that will be needed. A subset
S of a normed space is said to be totally bounded if for each ε > 0 there is a finite set
of open balls of radius ε which cover S. Clearly, a totally bounded set is bounded, i.e., it
is contained in a sufficiently large ball. It is not difficult to see that a relatively compact
subset of a normed space is totally bounded, with the converse being true if the normed
space is complete. Moreover, a totally bounded subset of a normed space is separable.
(a) The embedding W01,p (Ω) ⊂ Lq (Ω) is compact for each 1 ≤ q < np/(n − p).
(b) Assuming ∂Ω ∈ C 1 , the embedding W 1,p (Ω) ⊂ Lq (Ω) is compact for each 1 ≤
q < np/(n − p).
(c) Assuming ∂Ω ∈ C 1 , γ0 : W 1,p (Ω) → Lp (∂Ω) is compact.
If p > n, then
(d) Assuming ∂Ω ∈ C 1 , the embedding W 1,p (Ω) ⊂ C 0,α (Ω̄) is compact for each
0 ≤ α < 1 − (n/p).
Proof. We shall just give the proof for p = q = 2. The other cases are proved similarly. (a)
Since C01 (Ω) is dense in H01 (Ω), it suffices to show that the embedding C01 (Ω) ⊂ L2 (Ω) is
compact. Thus, let S = {u ∈ C01 (Ω) : kuk1,2 ≤ 1}. We now show that S is totally bounded
in L2 (Ω).
For h > 0, let Sh = {uh : u ∈ S}, where uh is the mollified function for u. We claim
that Sh is totally bounded in L2 (Ω). Indeed, for u ∈ S, we have
Z
|uh (x)| ≤ ωh (z)|u(x − z)|dz ≤ (sup ωh )kuk1 ≤ c1 (sup ωh )kuk1,2
B(0,h)
and
so that Sh is a bounded and equicontinuous subset of C(Ω̄). Thus by the Ascoli Theorem,
Sh is relatively compact (and thus totally bounded) in C(Ω̄) and consequently also in L2 (Ω).
Now, by earlier estimates, we easily obtain
Z Z
kuh − uk22 ≤ ωh (z) 2
|u(x − z) − u(x)| dx dz
B(0,h) Ω
46 2. Sobolev Spaces
and
1 2
du(x − tz)
Z Z Z
2
|u(x − z) − u(x)| dx = dt dx
Ω Ω 0 dt
Z Z 1 2
= (−∇u(x − tz) · z)dt dx
Ω 0
Z Z 1
2
≤ |z| |∇u(x − tz)| dt dx ≤ |z|2 kuk21,2 .
2
Ω 0
Consequently, kuh − uk2 ≤ h. Since we have shown above that Sh is totally bounded in
L2 (Ω) for all h > 0, it follows that S is also totally bounded in L2 (Ω) and hence relatively
compact.
(b) Suppose now that S is a bounded set in H 1 (Ω). Each u ∈ S has an extension
U ∈ H01 (Ω0 ) where Ω ⊂⊂ Ω0 . Denote by S 0 the set of all such extensions of the functions
u ∈ S. Since kU kH 1 (Ω0 ) ≤ ckuk1,2 , the set S 0 is bounded in H01 (Ω0 ). By (a) S 0 is relatively
compact in L2 (Ω0 ) and therefore S is relatively compact in L2 (Ω).
(c) Let S be a bounded set in H 1 (Ω). For any u(x) ∈ C 1 (Ω̄), the inequality (2.7) with
p = 2 yields
c1
(2.24) kuk2L2 (∂Ω) ≤ kuk22 + c2 βkuk21,2
β
where the constants c1 , c2 do not depend on u or β. By completion, this inequality is
valid for any u ∈ H 1 (Ω). By (b), any infinite sequence of elements of the set S has a
subsequence {un } which is Cauchy in L2 (Ω): given ε > 0, an N can be found such that for
all m, n ≥ N, kum − un k2 < ε. Now we choose β = ε. Applying the inequality (2.24) to
um − un , it follows that the sequence of traces {γ0 un } converges in L2 (∂Ω).
(d) By Morrey’s inequality, the embedding is continuous if α = 1 − (n/p). Now use the
fact that C 0,β is compact in C 0,α if α < β.
Remarks. (a) When p = n, we can easily show that the embedding in (a) is compact for
all 1 ≤ q < ∞. Hence, it follows that the embedding W01,p (Ω) ⊂ Lp (Ω) is compact for all
p ≥ 1. However, when p = n, we do not have embedding W 1,n (Ω) ⊂ L∞ (Ω). For example,
1
u = ln ln(1 + |x| ) ∈ W 1,n (B(0, 1)) but not to L∞ (B(0, 1)) if n ≥ 2.
(b) The boundedness of Ω is essential in the above theorem. For example, let I = (0, 1)
and Ij = (j, j + 1). Let f ∈ C01 (I) and define fj to be the same function defined on Ij by
translation. We can normalize f so that kf kW 1,p (I) = 1. The same is then true for each fj
and thus {fj } is a bounded sequence in W 1,p (R). Clearly f ∈ Lq (R) for every 1 ≤ q ≤ ∞.
Further, if
kf kLq (R) = kf kLq (I) = a > 0
then for any j 6= k we have
Z j+1 Z k+1
kfj − fk kqLq (R) = q
|fj | + |fk |q = 2aq
j k
and so fi cannot have a convergent subsequence in Lq (R). Thus none of the embeddings
W 1,p (R) ⊂ Lq (R) can be compact. This example generalizes to n dimensional space and to
open sets like a half-space.
2.5. Additional Topics 47
Proof. First of all, it is easy to check that k · k defines a norm. Now by (i), it suffices to
prove that there is a positive constant c such that
(2.25) kuk1,p ≤ ckuk for all u ∈ W 1,p (Ω).
Suppose (2.25) is false. Then there exist vn (x) ∈ W 1,p (Ω) such that kvn k1,p > nkvn k. Set
un = vn /kvn k1,p . So
(2.26) kun k1,p = 1 and 1 > nkun k.
According to Theorem 2.31, there is a subsequence, call it again {un }, which converges to u
in Lp (Ω). From (2.26) we have kun k → 0 and therefore ∇un → 0 in Lp (Ω) and q(un ) → 0.
From un → u, ∇un → 0 both in Lp (Ω), we have ∇u = 0 a.e. in Ω and hence u = C, a
constant. This also implies un → C in W 1,p (Ω) and, by kun k1,p = 1, C 6= 0. By continuity,
q(C) = 0, which implies C = 0 by (ii). We thus derive a contradiction.
Example 2.33. Let ∂Ω ∈ C 1 . Assume a(x) ∈ C(Ω), σ(x) ∈ C(∂Ω) with a ≥ 0 (6≡ 0), σ ≥
0 (6≡ 0). Then the following norms are equivalent to k · k1,p on W 1,p (Ω):
n p 1/p
Z X Z !
p
R
(2.27) kuk = |Di u| dx + udx with q(u) = Ω udx .
Ω i=1 Ω
n
!1/p
Z X Z p
|Di u|p dx +
R
(2.28) kuk = γ0 udS with q(u) = ∂Ω γ0 udS .
Ω i=1 ∂Ω
n
Z X Z !1/p
p p
R p
1/p
(2.29) kuk = |Di u| dx + σ|γ0 u| dS with q(u) = ∂Ω σ|γ0 u| dS .
Ω i=1 ∂Ω
48 2. Sobolev Spaces
n
Z X Z !1/p
p p
R p
1/p
(2.30) kuk = |Di u| dx + a|u| dx with q(u) = Ω a|u| dx .
Ω i=1 Ω
Clearly property (ii) of Theorem 2.32 is satisfied for these semi-norms q(u). In order to
verify condition (i), one uses the trace theorem in (2.28) and (2.29).
R
2.5.2. Poincaré’s Inequalities. Using u − (u)Ω , where (u)Ω = −Ω udx, in the equivalent
norm (2.27) we obtain that
Z n
Z X
p
(2.31) |u(x) − (u)Ω | dx ≤ c |Di u|p dx, u ∈ W 1,p (Ω)
Ω Ω i=1
where the constant c > 0 is independent of u. This is also called a Poincaré’s inequality.
Therefore !1/p
Z X n
p
kuk1,p,0 = |Di u| dx
Ω i=1
2.5.3. Difference Quotients. For later regularity theory, we will be forced to study the
difference quotient approximations to weak derivatives.
Assume u : Ω → RN is locally integrable. Let {e1 , · · · , en } be the standard basis of Rn .
Define the ith-difference quotient of size h of u by
u(x + hei ) − u(x)
Dih u(x) = , h 6= 0.
h
Then Dih u is defined on Ωh,i = {x ∈ Ω | x + hei ∈ Ω}. Note that
Ωh = {x ∈ Ω | dist(x; ∂Ω) > |h|} ⊂ Ωh,i .
We have the following properties of Dih u.
1) If u ∈ W 1,p (Ω; RN ) then Dih u ∈ W 1,p (Ωh,i ; RN ) and
D(Dih u) = Dih (Du) on Ωh,i .
(b) Let u ∈ Lp (Ω), 1 < p < ∞, and Ω0 ⊂⊂ Ω. If there exists a constant K > 0 such that
lim inf kDih ukLp (Ω0 ) ≤ K,
h→0
then the weak derivative Di u exists and satisfies kDi ukLp (Ω0 ) ≤ K.
Proof. (a) Let us suppose initially that u ∈ C 1 (Ω) ∩ W 1,p (Ω). Then, for h > 0,
1 h
Z
h
Di u(x) = Di u(x + thei ) dt,
h 0
so that by Hölder’s inequality
Z h
1
|Dih u(x)|p ≤ |Di u(x + thei )|p dt,
h 0
and hence Z Z hZ Z
1
|Dih u(x)|p dx ≤ p
|Di u| dx dt ≤ |Di u|p dx,
Ω0 h 0 Bh (Ω0 ) Ω
where Bh (Ω0 ) = {x ∈ Ω | dist(x; Ω0 ) < h}. The extension of this inequality to arbitrary
functions in W 1,p (Ω) follows by a straight-forward approximation argument.
(b) Since 1 < p < ∞, there exists a sequence {hm } tending to zero and a function
v ∈ Lp (Ω0 ) with kvkp ≤ K such that Dihm u * v in Lp (Ω0 ) as m → ∞. This means for all
φ ∈ C0∞ (Ω0 )
Z Z
lim φ Dihm u dx = φ v dx.
m→∞ Ω0 Ω0
Now for |hm | < dist(supp φ; ∂Ω0 ), we have
Z Z Z
φ Dihm u dx =− u Di−hm φ dx →− u Di φ dx.
Ω0 Ω0 Ω0
Hence Z Z
φ v dx = − u Di φ dx,
Ω0 Ω0
which shows v = Di u ∈ Lp (Ω0 ) and kDi ukLp (Ω0 ) ≤ K.
Remark 2.10. Variants of Theorem 2.34 can be valid even if it is not the case Ω0 ⊂⊂ Ω.
For example if Ω is the open half-ball B(0, 1) ∩ {xn > 0}, Ω0 = B(0, 1/2) ∩ {xn > 0}, and if
u ∈ W 1,p (Ω), then we have the bound
kDih ukLp (Ω0 ) ≤ kDi ukLp (Ω)
for i = 1, 2, · · · , n − 1 and 0 < |h| < dist(Ω0 , ∂Ω).
We will need this remark for boundary regularity later.
2.5.4. Fourier Transform Methods. For a function u ∈ L1 (Rn ), we define the Fourier
transform of u by
Z
1
û(y) = e−ix·y u(x) dx, ∀ y ∈ Rn ,
(2π)n/2 Rn
and the inverse Fourier transform by
Z
1
ǔ(y) = eix·y u(x) dx, ∀ y ∈ Rn .
(2π)n/2 Rn
50 2. Sobolev Spaces
Since L1 (Rn ) ∩ L2 (Rn ) is dense in L2 (Rn ), we can use this result to extend the Fourier
transforms on to L2 (Rn ). We still use the same notations for them. Then we have
(ii) Ddα u(y) = (iy)α û(y) for each multiindex α such that D α u ∈ L2 (Rn ),
ˇ
(iii) u = û.
Using the Fourier transform, we can also define fractional Sobolev spaces H s (Rn ) for
any 0 < s < ∞ as follows
where C = k(1 + |y|s )−2 k2L1 (Rn ) < ∞ if and only if s > n2 . Therefore we have an easy
embedding, which is known valid for integers s by the previous Sobolev embedding theorem,
2.6.1. Calculus of Abstract Functions. Let X be a real Banach space with norm k · k
and let I be any interval of the real line R.
Definition 2.11. (i) A function u : I → X is called an abstract function.
(ii) An abstract function u : I → X is said to be continuous at a point t0 ∈ I if
lim ku(t) − u(t0 )k = 0.
t→t0
(If t0 is an end point of I, the continuity at t0 is defined through the one-sided limit.) If
u(t) is continuous at each point of I, then we write u ∈ C(I; X).
(iii) Abstract function u : I → X is said to be differentiable at the point t0 ∈ I if
there exists an element l = u0 (t0 ) ∈ X such that
lim k[u(t0 + h) − u(t0 )]/h − u0 (t0 )k = 0.
h→0
We say u(t) is differentiable on I if it is differentiable at each point of I.
Remark 2.12. (i) If u : I → X is continuous at t0 ∈ I then real-valued function ku(t)k is
continuous at t0 .
(ii) If I = [a, b] is a compact interval, then C([a, b]; X) becomes a Banach space with
norm
kukC([a,b];X) := max ku(t)k.
t∈[a,b]
Theorem 2.38. (Mean Value Theorem) Let u(t) ∈ C([a, b]; X) and suppose u0 (t) exists
for every t ∈ (a, b). Then
ku(a) − u(b)k ≤ (b − a) sup ku0 (t)k.
a<t<b
Proof. We use a standard device which reduces the problem to the classical case. Namely,
consider the real-valued function φ(t) = f (u(t)), where f ∈ X ∗ . Since f is continuous and
linear we have
0 u(t + h) − u(t)
φ (t) = lim f = f (u0 (t)).
h→0 h
Now apply the classical mean value theorem to φ(·) to get
ku(b) − u(a)k = sup f (u(b) − u(a))
kf k=1
= sup (φ(b) − φ(a))
kf k=1
if such a common limit value exists for all sequences of partitions {Zk } for which ∆Zk → 0
as k → ∞.
Theorem 2.39. If u(t) ∈ C([a, b]; X), then the Riemann integral exists.
Proof. As in the classical proof, one uses the uniform continuity of u(t) together with the
completeness of X. We shall omit the details.
Theorem 2.40. Let u(t) : [a, b] → X be continuous. Then the following hold:
(a)
Z b Z b
k u(t)dtk ≤ ku(t)kdt
a a
(b)
Z b Z b
f u(t)dt = f (u(t))dt for all f ∈ X∗
a a
(c)
d t
Z
u(s)ds = u(t) for all a ≤ t ≤ b
dt a
(d) If u0 (t) ∈ C((a, b); X), then for any α, β ∈ (a, b)
Z β
u0 (s)ds = u(β) − u(α)
α
Proof. (a) and (b) follow by passing to the limit in the corresponding relations of Riemann
sums.
Rt
(c) Set v(t) = a u(s)ds. Since u(t) is uniformly continuous on [a, b], we have
Z t+h
−1
k[v(t + h) − v(t)]/h − u(t)k = kh [u(s) − u(t)]dsk
t
≤ max ku(s) − u(t)k → 0 as h → 0.
|s−t|≤|h|
2.6.2. Measurable Functions and Sobolev Spaces. We now extend the continuous
functions to measurable functions. In what follows, we assume I is a bounded interval in R
and X is a Banach space.
Definition 2.14. (i) A function s : I → X is called simple if it has the form
m
X
s(t) = χEi (t)ui (t ∈ I),
i=1
where Ei is a Lebesgue measurable subset of I and ui ∈ X for i = 1, 2, · · · , m. In this case,
we define
Z m
X
s(t) dt = |Ei |ui ∈ X.
I i=1
2.6. Spaces of Functions Involving Time 53
Remark 2.15. (i) The space Lp (I; X) consists of all strongly measurable functions
u : I → X with
Z 1/p
kukLp (I;X) := ku(t)kp dt <∞
I
if 1 ≤ p < ∞ and
kukL∞ (I;X) := esssupt∈I ku(t)k < ∞
if p = ∞. As in the usual Lebesgue space cases, we identify functions that are almost
everywhere equal. Then Lp (I; X) becomes a Banach space for all 1 ≤ p ≤ ∞.
(ii) If X is reflexive, then we have
p
(Lp (I; X))∗ ≈ Lq (I; X ∗ ) (1 < p < ∞, q = ).
p−1
However, usually, (L1 (I; X))∗ 6≈ L∞ (I; X ∗ ). In fact, if X is a separable Banach space, then
(L1 (I; X))∗ ≈ L∞ ∗
w (I; X ),
where L∞ ∗ ∗
w (I; X ) consists of functions g : I → X such that for each u ∈ X the function
t 7→ hg(t), ui is Lebesgue measurable and essentially bounded on I with the norm
kgkw := sup khg(t), uikL∞ (I) < ∞.
u∈X,kuk≤1
Definition 2.16. (i) Let u, v ∈ L1 (I; X). We say v is the weak derivative of u, written
u0 = v, provided Z Z
0
φ (t)u(t) dt = − φ(t)v(t) dt
I I
holds in X for all scalar test functions φ ∈ Cc∞ (I).
(ii) The Sobolev space W 1,p (I; X) consists of all functions u ∈ Lp (I; X) such that weak
derivative u0 exists and belongs to Lp (I; X). The norm is defined by
( R 1/p
I (ku(t)kp + ku0 (t)kp )dt (1 ≤ p < ∞),
kukW 1,p (I;X) = 0
esssupt∈I (ku(t)k + ku (t)k) (p = ∞).
We write H 1 (I; X) = W 1,2 (I; X).
Theorem 2.42. (Calculus in W 1,p (I; X)) Let u ∈ W 1,p (I; X) for some 1 ≤ p ≤ ∞. Then
¯ X) (after being redefined on a null set of time), with
(i) u ∈ C(I;
kukC(I;X)
¯ ≤ CkukW 1,p (I;X)
for a constant C depending on I.
Rt
(ii) u(t) = u(s) + s u0 (τ ) dτ for all s ≤ t in I.
For use later in the regularity study, we will need an extension of Theorem 2.43.
Theorem 2.44. (Mapping into better spaces) Let Ω be a bounded domain with smooth
∂Ω and m a nonnegative integer. Suppose u ∈ L2 (I; H m+2 (Ω)), with u0 ∈ L2 (I; H m (Ω)).
¯ H m+1 (Ω)) (after being redefined on a null set of time), with
Then u ∈ C(I;
0
kukC(I;H
¯ m+1 (Ω)) ≤ C(kukL2 (I;H m+2 (Ω)) + ku kL2 (I;H m (Ω)) )
Second-Order Linear
Elliptic Equations
3.1.1. Linear Elliptic Equations. We study the (Dirichlet) boundary value problem
(BVP)
(3.1) Lu = f in Ω, u = 0 on ∂Ω.
Here f is a given function in L2 (Ω) (or more generally, an element in the dual space of
H01 (Ω)) and L is a second-order differential operator having either the divergence
form
Xn n
X
(3.2) Lu ≡ − Di (aij (x)Dj u) + bi (x)Di u + c(x)u
i,j=1 i=1
or else
n
X n
X
Lu ≡ − aij (x)Dij u + bi (x)Di u + c(x)u
i,j=1 i=1
with given real coefficients aij (x), bi (x) and c(x). We also assume
aij (x) = aji (x) (i, j = 1, . . . , n).
Definition 3.1. The partial differential operator L is said to be uniformly elliptic in
Ω if there exists a number θ > 0 such that for every x ∈ Ω and every real vector ξ =
(ξ1 , . . . , ξn ) ∈ Rn
n
X n
X
(3.3) aij (x)ξi ξj ≥ θ |ξi |2 .
i,j=1 i=1
55
56 3. Second-Order Linear Elliptic Equations
Definition 3.2. Let f ∈ L2 (Ω). A function u ∈ H01 (Ω) is called a weak solution of (3.1)
with L given by (3.2) if B1 [u, v] = (f, v)L2 for all v ∈ H01 (Ω); that is, the following holds
Z n n Z
X X
f vdx ∀ v ∈ H01 (Ω).
(3.4) aij Dj uDi v + ( bi Di u + cu)v dx =
Ω i,j=1 i=1 Ω
If f ∈ H −1 (Ω) = (H01 (Ω))∗ , the dual space of H01 (Ω), then weak solutions are defined by
replacing the right-hand side by hf, ui, h, i being the dual pairing on H −1 (Ω) × H01 (Ω).
Exercise 3.3. Consider the following weak formulation: Given f ∈ L2 (Ω). Find u ∈ H 1 (Ω)
satisfying
Z Z
∇u · ∇vdx = f vdx for all v ∈ H 1 (Ω).
Ω Ω
Find the boundary value problem solved by u. What is the necessary condition for the
existence of such a u?
The (Dirichlet) BVP for a most general system of second-order (quasilinear) par-
tial differential equations in divergence form can be written as follows:
where A(x, s, ξ) = (Aki (x, u, ξ)), 1 ≤ i ≤ n, 1 ≤ k ≤ N, and b(x, s, ξ) = (bk (x, u, ξ)),
1 ≤ k ≤ N, are given functions of (x, u, ξ) ∈ Ω × RN × MN ×n , and F = (f k ), 1 ≤ k ≤ N ,
with each f k being a given functional in the dual space of W01,p (Ω).
The coefficients A, b usually satisfy certain structural conditions that will generally
0
assure that both |A(x, u, Du)| and |b(x, u, Du)| belong to Lp (Ω) for all u ∈ W 1,p (Ω; RN ),
where p0 = p−1 p
. In such cases, a function u ∈ W01,p (Ω; RN ) is called a weak solution of
(3.5) if the following holds
Z "X n
#
(3.6) Ai (x, u, Du)Di ϕ + b (x, u, Du)ϕ dx = hf k , ϕi
k k
Ω i=1
Definition 3.4. The system (3.5) is said to be linear if both A and b are linear in the
variables (u, ξ); that is,
X N
X
Aki (x, u, Du) = akl
ij (x) Dj u
l
+ dkl l
i (x) u ,
1≤l≤N, 1≤j≤n l=1
(3.7)
X N
X
k
b (x, u, Du) = bkl
j (x) Dj u
l
+ ckl (x) ul .
1≤j≤n, 1≤l≤N l=1
3.1. Differential Equations in Divergence Form 57
For linear systems, the suitable space is Hilbert space H01 (Ω; RN ) equipped with the
inner product defined by
X Z
(u, v) ≡ Di uk Di v k dx.
1≤i≤n, 1≤k≤N Ω
Ellipticity Conditions. There are several ellipticity conditions for the system (3.5) in
terms of the leading coefficients A(x, u, ξ). Assume A is smooth on ξ and define
∂Aki (x, u, ξ)
Akl
ij (x, u, ξ) = , ξ = (ξjl ).
∂ξjl
The system (3.5) is said to satisfy the (uniform, strict) Legendre ellipticity condition
if there exists a ν > 0 such that, for all (x, s, ξ), it holds
n X
X N
(3.9) Akl k l
ij (x, s, ξ) ηi ηj ≥ ν |η|
2
for all N × n matrix η = (ηik ).
i,j=1 k,l=1
For systems with linear leading terms A given by (3.7), the Legendre condition and
Legendre-Hadamard condition become, respectively,
n X
X N
(3.11) akl k l
ij (x) ηi ηj ≥ ν |η|
2
∀ η;
i,j=1 k,l=1
n X
X N
(3.12) akl k l 2
ij (x) q q pi pj ≥ ν |p| |q|
2
∀ p, q.
i,j=1 k,l=1
Exercise 3.5. If N > 1, the Legendre-Hadamard condition does not imply the Legendre
ellipticity condition. For example, let n = N = 2 and ε > 0. Define constants akl
ij by
2
X
akl k l 2
ij ξi ξj ≡ det ξ + ε |ξ| .
i,j,k,l=1
58 3. Second-Order Linear Elliptic Equations
Show that the Legendre-Hadamard condition holds for all ε > 0. But, the Legendre condi-
tion holds for this system if and only if ε > 1/2.
Exercise 3.6. Let u = (v, w) and x = (x1 , x2 ) = (x, y) ∈ R2 . Then the system of differential
equations defined by akl
ij given above is
∆v + wxy = 0, ∆w − vxy = 0.
This system reduces to two fourth-order equations for v, w (where ∆f = fxx + fyy ):
2 ∆2 v − vxxyy = 0, 2 ∆2 w + wxxyy = 0.
We can easily see that both equations are elliptic if and only if > 1/2.
Exercise 3.7. Formulate the biharmonic equation ∆2 u = f as a linear system and find the
appropriate bilinear form B[u, v] in the definition of weak solutions.
Proof. For each fixed u ∈ H, the functional v 7→ B[u, v] is in H ∗ , and hence by the Riesz
Representation Theorem, there exists a unique element w = Au ∈ H such that
B[u, v] = (w, v) ∀ v ∈ H.
It can be easily shown that A : H → H is linear. From (i), kAuk2 = B[u, Au] ≤ αkukkAuk,
and hence kAuk ≤ αkuk for all u ∈ H; that is, A is bounded. Furthermore, by (ii),
βkuk2 ≤ B[u, u] = (Au, u) ≤ kAukkuk and hence kAuk ≥ βkuk for all u ∈ H. By the
Riesz Representation Theorem again, we have a unique w0 ∈ H such that hf, vi = (w0 , v)
for all v ∈ H and kf k = kw0 k. We will show that the equation Au = w0 has a (unique)
solution. There are many different proofs for this, and here we use the Contraction Mapping
Theorem. Note that the solution u to equation Au = w0 is equivalent to the fixed-point
of the map T : H → H defined by T (v) = v − tAv + tw0 (v ∈ H) for any fixed t > 0. We
will show for t > 0 small enough T is a contraction. Note that for all v, w ∈ H we have
kT (v) − T (w)k = k(I − tA)(v − w)k. We compute that for all u ∈ H
k(I − tA)uk2 = kuk2 + t2 kAuk2 − 2t(Au, u)
≤ kuk2 (1 + t2 α2 − 2βt).
3.3. Energy Estimates and Existence Theory 59
We now choose t such that 0 < t < α2β2 . Then the expression in parentheses is positive and
less than 1. Thus the map T : H → H is a contraction on H and therefore has a fixed point.
This fixed point u solves Au = w0 and thus is the unique solution of (3.13); moreover, we
have kf k = kw0 k = kAuk ≥ βkuk and hence kuk ≤ β1 kf k. The proof is complete.
Proof. Take γ from (3.14), let λ ≥ γ and define the bilinear form
B λ [u, v] ≡ B1 [u, v] + λ(u, v)2 for all u, v ∈ H
which corresponds to the operator Lu + λu. Then B λ [u, v] satisfies the hypotheses of the
Lax-Milgram Theorem.
Example 3.4. Consider the Neumann boundary value problem
∂u
(3.15) −∆u(x) = f (x) in Ω, = 0 on ∂Ω.
∂ν
A function u ∈ H 1 (Ω) is said to be a weak solution to (3.15 if
Z Z
(3.16) ∇u · ∇v dx = f v dx, ∀ v ∈ H 1 (Ω).
Ω Ω
Obviously,
R taking v ≡ 1 ∈ H 1 (Ω), a necessary condition to have a weak solution is
Ω f (x) dx = 0. We show this is also a sufficient condition for existence of the weak so-
lutions. Note that, if u is a weak solution, then u + c, for all constants c, is also a weak
solution. Therefore, to fix the constants, we consider the vector space
Z
1
H = u ∈ H (Ω) u(x) dx = 0
Ω
equipped with inner product
Z
(u, v)H = ∇u · ∇v dx.
Ω
By the theorem on equivalent norms, it follows that H with this inner product, is indeed a
Hilbert space, and (f, u)L2 (Ω) is a bounded linear functional on H:
|(f, u)L2 (Ω) | ≤ kf kL2 (Ω) kvkL2 (Ω) ≤ kf kL2 (Ω) kvkH .
Hence the Riesz Representation Theorem implies that there exists a unique u ∈ H such
that
(3.17) (u, w)H = (f, w)L2 (Ω) , ∀ w ∈ H.
It follows that u is a weak solutionR to the Neumann problem since for any Rv ∈ H 1 (Ω) we
1
take w = v − c ∈ H, where c = |Ω| Ω vdx, in (3.17) and obtain (3.16) using Ω f dx = 0.
Example 3.5. Let us consider the nonhomogeneous Dirichlet boundary value problem
(3.18) −∆u = f in Ω, u|∂Ω = ϕ
where f ∈ L2 (Ω) and ϕ is the trace of a function w ∈ H 1 (Ω). Note that it is not sufficient
to just require that ϕ ∈ L2 (∂Ω) since the trace operator is not onto. If, for example,
ϕ ∈ C 1 (∂Ω), then ϕ has a C 1 extension to Ω̄, which is the desired w.
The function u ∈ H 1 (Ω) is called a weak solution of (3.18) if u − w ∈ H01 (Ω) and if
Z Z
∇u · ∇vdx = f vdx for all v ∈ H01 (Ω).
Ω Ω
Let u be a weak solution of (3.18) and set u = z + w. Then z ∈ H01 (Ω) satisfies
Z Z
(3.19) ∇z · ∇vdx = (f v − ∇v · ∇w)dx for all v ∈ H01 (Ω).
Ω Ω
Since the right hand side belongs to the dual space H −1 (Ω) = H01 (Ω)∗ , the Lax-Milgram
theorem yields the existence of a unique z ∈ H01 (Ω) which satisfies (3.19). Hence (3.18) has
a unique weak solution u.
3.3. Energy Estimates and Existence Theory 61
Example 3.6. Now let us consider the boundary value (also called Dirichlet) problem for
the fourth order biharmonic operator:
∂u
∆2 u = f in Ω, u|∂Ω = |∂Ω = 0.
∂n
We take H = H02 (Ω). By the general trace theorem, H = H02 (Ω) = {v ∈ H 2 (Ω) : γ0 v =
γ1 v = 0}. Therefore, this space H is the right space for the boundary conditions.
Accordingly, for f ∈ L2 (Ω), a function u ∈ H = H02 (Ω) is a weak solution of the
Dirichlet problem for the biharmonic operator provided
Z Z
∆u∆vdx = f vdx ∀ v ∈ H.
Ω Ω
Consider the bilinear form Z
B[u, v] = ∆u∆vdx.
Ω
Its boundedness follows from the Cauchy-Schwarz inequality
|B[u, v]| ≤ k∆uk2 k∆vk2 ≤ dkuk2,2 kvk2,2 .
Furthermore, it can be shown that k∆uk2 defines a norm on H02 (Ω) which is equivalent to
the usual norm on H 2 (Ω). (Exercise!) Hence
B[u, u] = k∆uk22 ≥ ckuk22,2
and so, by the Lax-Milgram theorem (in fact, just the Riesz Representation Theorem), there
exists a unique weak solution u ∈ H.
Exercise 3.8. Denote by Hc1 the space
Hc1 = {u ∈ H 1 (Ω) : γ0 u = const}.
Note that the constant may be different for different u’s.
(a) Prove that Hc1 is complete.
(b) Let f ∈ C(Ω̄). Prove existence of a unique u ∈ Hc1 satisfying
Z Z
(∇u · ∇v + uv)dx = f vdx ∀ v ∈ Hc1 .
Ω Ω
(c) If u ∈ C 2 (Ω̄) satisfies the equation in (b), find the underlying BVP.
Exercise 3.9. Let Ω = (1, +∞). Show that the BVP −u00 = f ∈ L2 (Ω), u ∈ H01 (Ω) does
not have a weak solution.
3.3.2. Gårding’s estimate for B2 [u, u]. We will derive the Gårding estimate for B2 [u, u].
For simplicity, let H = H01 (Ω; RN ) and let (u, v)H and kukH be the equivalent inner product
and norm defined above on H. Define the bilinear form of the leading terms by
Xn X N Z
A[u, v] = akl l k
ij (x) Dj u Di v dx.
i,j=1 k,l=1 Ω
Proof. In the first case, the conclusion follows easily from the Legendre condition. We
prove the second case when Aklij are constants satisfying the Legendre-Hadamard condition
n X
X N
akl k l 2 2
ij q q pi pj ≥ ν |p| |q| , ∀p ∈ Rn , q ∈ RN .
i,j=1 k,l=1
We prove
n X
X N Z Z
A[u, u] = akl l k
ij Dj u Di u dx ≥ ν |Du|2 dx
i,j=1 k,l=1 Ω Ω
for all u ∈ C0∞ (Ω; RN ). For these test functions u, we extend them onto Rn by zero outside
Ω and thus consider them as functions in C0∞ (Rn ; RN ). Define the Fourier transforms
for such functions u by
Z
−n/2
û(y) = (2π) e−i y·x u(x) dx; y ∈ Rn .
Rn
D[ k ck
j u (y) = i yj u (y);
the last identity can also be written as Du(y) c = i û(y) ⊗ y. Now, using these identities, we
have Z Z
kl k l
aij Di u (x) Dj u (x) dx = akl [k [l
ij Di u (y) Dj u (y) dy
Rn R n
Z Z
= akl y y
ij i j
ck (y) ubl (y) dy = Re
u akl
y y
ij i j
ck (y) ubl (y) dy .
u
Rn Rn
Write û(y) = η + iξ with η, ξ ∈ RN . Then
Re u (y) u (y) = η k η l + ξ k ξ l .
ck bl
i,j=1 k,l=1
Hence,
n X
X N Z
A(u, u) = akl k l
ij Di u (x) Dj u (x) dx
i,j=1 k,l=1 Rn
n
X N Z
X
= Re akl
ij yi yj
ck (y) ubl (y) dy
u
i,j=1 k,l=1 Rn
Z Z
2 2
≥ν |y| |û(y)| dy = ν |iû(y) ⊗ y|2 dy
Rn Rn
Z Z
2
=ν |Du(y)|
c dy = ν |Du(x)|2 dx.
Rn Rn
The proof is complete.
3.3. Energy Estimates and Existence Theory 63
Theorem 3.8. (Gårding’s estimate for system) Let B2 [u, v] be defined by (3.8). Assume
1) akl
ij ∈ C(Ω̄),
2) the Legendre-Hadamard condition holds for all x ∈ Ω; that is,
akl k l 2 2
ij (x) q q pi pj ≥ ν |p| |q| , ∀p ∈ Rn , q ∈ RN .
3) bkl kl kl ∞
i , c , di ∈ L (Ω).
Then, there exist constants λ0 > 0 and λ1 ≥ 0 such that
B2 [u, u] ≥ λ0 kuk2H − λ1 kuk2L2 , ∀ u ∈ H01 (Ω; RN ).
for all test functions u ∈ C0∞ (Ω; RN ) with diam(supp u) ≤ . To see this, we choose any
point x0 ∈ supp u. Then
Z Z
kl k l
aij (x) Di u Dj u dx = akl k l
ij (x0 ) Di u Dj u dx
Ω Ω
Z
akl kl k l
+ ij (x) − aij (x0 ) Di u Dj u dx
supp u
Z Z
ν 2
≥ν |Du(x)| dx − |Du(x)|2 dx,
Ω 2 Ω
which proves (3.20).
2. Now assume u ∈ C0∞ (Ω; RN ), with arbitrary compact support. We cover Ω̄ with
finitely many open balls {B/4 (xm )} with xm ∈ Ω and m = 1, 2, ..., M. For each m, let
ζm ∈ C0∞ (B/2 (xm )) with ζm (x) = 1 for x ∈ B/4 (xm ). Since for any x ∈ Ω̄ we have at least
one m such that x ∈ B/4 (xm ) and thus ζm (x) = 1, we may therefore define
ζm (x)
ϕm (x) = PM 2 1/2 , m = 1, 2, ..., M.
j=1 ζj (x)
Then M 2
P
m=1 ϕm (x) = 1 for all x ∈ Ω. (This is a special case of partition of unity.) We
have thus
M
X
akl k l
ij (x) Di u Dj u = akl 2 k
ij (x) ϕm Di u Dj u
l
m=1
M
X
(3.21) = akl k l
ij (x) Di (ϕm u ) Dj (ϕm u )
m=1
M
X
− akl
ij (x) ϕm ul
D ϕ D
i m i uk
+ ϕm uk
D ϕ D
i m j ul
+ uk l
u Di m j m .
ϕ D ϕ
m=1
64 3. Second-Order Linear Elliptic Equations
Since ϕm u ∈ C0∞ (Ω ∩ B/2 (xm ); RN ) and diam(Ω ∩ B/2 (xm )) ≤ , we have by (3.20)
Z Z
kl k l ν X
aij (x) Di (ϕm u )Dj (ϕm u ) dx ≥ |Di (ϕm uk )|2 dx
Ω 2 1≤i≤n Ω
1≤k≤N
Z
ν X
= ϕ2m |Di uk |2 + |Di ϕm |2 |uk |2 + 2ϕm uk Di ϕm Di uk dx
2 1≤i≤n Ω
1≤k≤N
Z
ν X
≥ ϕ2m |Di uk |2 dx + 2ϕm uk Di ϕm Di uk dx
2 1≤i≤n Ω
1≤k≤N
Z Z
ν X 2 k 2 2 ν
≥ ϕ |Di u | dx − CkukL2 (Ω) = ϕ2 |Du|2 dx − Ckuk2L2 (Ω) ,
4 1≤i≤n Ω m 4 Ω m
1≤k≤N
where we have used the Cauchy inequality with . Then by (3.21) and the fact that
PM 2
m=1 ϕm = 1 on Ω,
Z Z
kl k l ν
aij (x) Di u Dj u dx ≥ |Du|2 dx − CM kuk2L2 (Ω) − C1 kukL2 (Ω) kDukL2 (Ω) .
Ω 4 Ω
The terms in B2 [u, u] involving b, c and d can all be estimated by
C2 (kukL2 (Ω) kDukL2 (Ω) + kuk2L2 (Ω) ).
Finally, using the Cauchy inequality with again, we have
ν
B2 [u, u] ≥ kuk2H 1 (Ω) − C3 kuk2L2 (Ω) ∀ u ∈ C0∞ (Ω; RN )
8 0
and, by density, for all u ∈ H01 (Ω; RN ). This completes the proof.
Note that the bilinear form B λ [u, v] = B2 [u, v] + λ (u, v)L2 satisfies the condition of
the Lax-Milgram theorem on H = H01 (Ω; RN ) for all λ ≥ λ1 ; thus, by the Lax-Milgram
theorem, we easily obtain the following existence result.
Theorem 3.9. Under the hypotheses of the previous theorem, for λ ≥ λ1 , the Dirichlet
problem for the system (3.5) with linear coefficients (3.7) has a unique weak solution u in
H01 (Ω; RN ) for any bounded linear functional F on H. Moreover, the solution u satisfies
kukH ≤ C kF k with a constant C depending on λ, and the coefficients.
Corollary 3.10. Given λ ≥ λ1 as in the theorem, then the operator K : L2 (Ω; RN ) →
L2 (Ω; RN ), where, for each F ∈ L2 (Ω; RN ), u = KF is the unique weak solution to the
BVP above, is a compact linear operator.
Proof. By the theorem, kukH01 (Ω;RN ) ≤ CkF kL2 (Ω;RN ) . Hence K is a bounded linear opera-
tor from L2 (Ω; RN ) to H01 (Ω; RN ), which, by the compact embedding theorem, is compactly
embedded in L2 (Ω; RN ). Hence, as a linear operator from L2 (Ω; RN ) to L2 (Ω; RN ), K is
compact.
X N
X
Ãki (x, u, Du) = ãkl l
ij (x) Dj u + d˜kl l
i (x) u ,
1≤l≤N, 1≤j≤n l=1
(3.23)
X N
X
b̃k (x, u, Du) = b̃kl l
j (x) Dj u + c̃kl (x) u,
1≤j≤n, 1≤l≤N l=1
where
ãkl lk
ij = aji , d˜kl lk
i = bi , b̃kl lk
j = di , c̃kl = clk (1 ≤ i, j ≤ n, 1 ≤ k, l ≤ N ).
The ellipticity condition of L∗ u is the same as that of Lu, and also B2∗ [u, u] = B2 [u, u].
Theorem 3.11. (Second Existence Theorem for weak solutions) Assume the ellip-
ticity and boundedness of the coefficients of Lu.
(i) Precisely one of the following statements holds:
either
(
for each F ∈ L2 (Ω; RN ) there exists a unique
(3.24)
weak solution u ∈ H01 (Ω; RN ) of Lu = F,
or else
(3.25) there exists a weak solution u 6= 0 in H01 (Ω; RN ) of Lu = 0.
(ii) Furthermore, should case (3.25) hold, the dimension of the subspace N ⊂ H01 (Ω; RN )
of weak solutions of Lu = 0 is finite and equals the dimension of the subspace
N ∗ ⊂ H01 (Ω; RN ) of weak solutions of adjoint problem L∗ u = 0.
(iii) Finally, the problem Lu = F has a weak solution if and only if
(F, v)L2 (Ω;RN ) = 0 ∀ v ∈ N ∗.
The dichotomy (3.24), (3.25) is called the Fredholm alternatives.
Proof. We assume the Gårding inequality holds (see Theorem 3.8 for sufficient conditions):
(3.26) B2 [u, u] ≥ σkuk2H 1 − µkuk2L2 , ∀ u ∈ H01 (Ω; RN ),
0
where σ > 0 and µ ∈ R are constants. We also assume µ > 0. For each F ∈ L2 (Ω; RN ),
define u = KF to be the unique weak solution in H01 (Ω; RN ) of the BVP
Lu + µu = F in Ω, u|∂Ω = 0.
By Theorem 3.9 and Corollary 3.10, this K is well defined and is a compact linear oper-
ator on L2 (Ω; RN ). We write K = (L + µI)−1 . Here I denotes the identity on L2 (Ω; RN )
and also the identity embedding of H01 (Ω; RN ) into L2 (Ω; RN ). Furthermore, given F ∈
L2 (Ω; RN ), u ∈ H01 (Ω; RN ) is a weak solution of Lu = F if and only if Lu + µu = F + µu,
which is equivalent to the equation u = K(F + µu) = KF + µKu; that is, (I − µK)u = KF.
Hence, we have N = N (I − µK) and similarly, N ∗ = N (I − µK∗ ); moreover, Lu = F if
66 3. Second-Order Linear Elliptic Equations
and only if KF ∈ R(I − µK) = (N (I − µK∗ ))⊥ = (N ∗ )⊥ . The proof of (iii) follows as, for
all v ∈ N ∗ , v = µK∗ v and so
(F, v) = (F, µK∗ v) = µ(KF, v);
hence KF ∈ (N ∗ )⊥ if and only if F ∈ (N ∗ )⊥ .
Here akl ∞
ij (x) and c(x) are given functions in L (Ω). The bilinear form associated to L is
Z XN X n
(3.27) B[u, v] = akl l k
ij (x)Dj u Di v + c(x)u · v
dx, u, v ∈ H01 (Ω; RN ).
Ω k,l=1 i,j=1
We also assume the Gårding inequality holds (see Theorem 3.8 for sufficient conditions):
(3.29) B[u, u] ≥ σkuk2H 1 − µkuk2L2 , ∀ u ∈ H01 (Ω; RN ),
0
−µ < λ1 ≤ λ2 ≤ λ3 ≤ · · ·
are listed repeatedly the same times as the multiplicity, and
lim λk = ∞.
k→∞
Proof. 1. From Theorem 3.12, we see that λ is an eigenvalue of L if and only if equation
(I − (λ + µ)K)u = 0 has nontrivial solutions u ∈ L2 (Ω; RN ); this exactly says that λ 6= −µ
1
and λ+µ is an eigenvalue of operator K. Since, by (3.30), K is strictly positive, all eigenvalues
of K consist of a countable set of positive numbers tending to zero and hence the eigenvalues
of L consist of a set of numbers {λj }∞j=1 with
−µ < λ1 ≤ λ2 ≤ · · · ≤ λj → ∞.
Set w̃k = (λk + µ)−1/2 wk , and consider the inner product on H = H01 (Ω; RN ) defined by
((u, v)) := Bµ [u, v] = B[u, v] + µ(u, v)L2 (Ω) (u, v ∈ H).
68 3. Second-Order Linear Elliptic Equations
This implies ((um , um )) ≤ ((u, u)) for all m = 1, 2, · · · . Hence, {um } is bounded in H and
so, by a subsequence, um * ũ in H as m → ∞. Since um → u in L2 , we must have ũ = u
and so
((u, u)) ≤ lim inf ((um , um )),
m→∞
which, combined with ((u − um , u − um )) = ((u, u)) + ((um , um )) − 2((u, um )) = ((u, u)) −
((um , um )), implies that um → u in H, and the claim is proved.
3. Now, by (3.32), we have
∞
X ∞
X ∞
X
B[u, u] = dk B[wk , u] = d2k λk ≥ d2k λ1 = λ1 .
k=1 k=1 k=1
where the uniform ellipticity condition is satisfied, ∂Ω is smooth, and aij , c are smooth
functions satisfying
aij (x) = aji (x), c(x) ≥ 0 (x ∈ Ω̄).
Theorem 3.14. The principal eigenvalue λ1 > 0. Let w1 be an eigenfunction corresponding
to the principal eigenvalue λ1 of L above. Then, either w1 (x) > 0 for all x ∈ Ω or w1 (x) < 0
for all x ∈ Ω. Moreover, the eigenspace corresponding to λ1 is one-dimensional.
Proof. 1. Since in this case the bilinear form B is positive: B[u, u] ≥ σkuk2H 1 (Ω) , we have
0
λ1 > 0. Let w1 be an eigenfunction corresponding to λ1 with kw1 kL2 (Ω) = 1, and set
w1+ = max{0, w1 }, w1− = min{0, w1 }.
3.5. Regularity 69
Then w1± ∈ H01 (Ω), w1 = w1+ + w1− , kw1+ k2L2 (Ω) + kw1− k2L2 (Ω) = kw1 k2L2 (Ω) = 1, and
2. Since the coefficients of L and Ω are smooth, u = w1± are smooth solutions. (See
Theorem 3.19 below.) Note that Lw1+ = λ1 w1+ ≥ 0 in Ω. By Strong Maximum Principle,
either w1+ ≡ 0 or else w1+ > 0 in Ω; similarly, either w1− ≡ 0 or else w1− < 0 in Ω. This
proves that either w1 < 0 in Ω or else w1 > 0 in Ω.
3. We now prove the eigenspace of λ1 is one-dimensional. Let w be another eigenfunc-
tion. Then, either w(x) > 0 for all x ∈ Ω or w(x) < 0 for all x ∈ Ω. Let t ∈ R be such
that Z Z
w(x) dx = t w1 (x)dx.
Ω Ω
Note that u = w − tw1 is also a solution to Lu = λ1 u. We claim u ≡ 0 and hence w = tw1 ,
proving the eigenspace is one-dimensional. Suppose u 6≡ 0. Then u is another eigenfunction
corresponding to λ1 . Then, by the theorem, weR would have either u(x) > 0 for all x ∈ Ω or
u(x) < 0 for all x ∈ Ω; hence, in either case, Ω u(x)dx 6= 0, which is a contradiction.
Remark 3.11. Let L = −∆. Then, there exists an orthonormal basis {wk }∞ k=1 of L (Ω)
2
1
consisting eigenfunctions wk of −∆ in H0 (Ω). We can see that {wk } is also orthogonal in
H01 (Ω); in fact,
Z
∇wk (x) · ∇wl (x) dx = B[wk , wl ] = λk (wk , wl )L2 (Ω) = λk δkl (k, l = 1, 2, · · · ).
Ω
3.5. Regularity
We now address the question as to whether a weak solution u of the PDE
Lu = f in Ω
is smooth or not. This is the regularity problem for weak solutions.
We first study second-order linear differential equations of the divergence form
n
X n
X
(3.33) Lu ≡ − Di (aij (x)Dj u) + bi (x)Di u + c(x)u
i,j=1 i=1
70 3. Second-Order Linear Elliptic Equations
To see that there is some hope that a weak solution may be better than a typical function
in H01 (Ω), let us consider the model problem −∆u = f in Rn . Assume u is smooth enough
to justify the following calculations.
Z Z Xn Z
2 2
f dx = (∆u) dx = uxi xi uxj xj dx
Rn Rn i,j=1 R
n
Xn Z
= − uxi xi xj uxj dx
n
i,j=1 R
n
X Z Z
= uxi xj uxi xj dx = |D2 u|2 dx.
i,j=1 Rn Rn
Thus the L2 norm of the second derivatives of u can be estimated by the L2 norm of f .
Similarly, if we differentiate the PDE with respect to xk , we see that the L2 norm of the
third derivatives of u can be estimated by the L2 norm of the first derivatives of f , etc. This
suggests that we can expect a weak solution u ∈ H01 (Ω) to belong to H m+2 (Ω) whenever
f ∈ H m (Ω).
The above calculations do not really constitute a proof, since we assumed that u was
smooth in order to carry out the calculation. If we merely start with a weak solution in
H01 (Ω), we cannot justify the above computations.
3.5.1. Difference Quotient Method. One can instead rely upon an analysis of certain
difference quotients to obatin higher regularity of weak solutions in H 1 (Ω). Our first regu-
larity result provides the interior H 2 -regularity for weak solutions of the equation Lu = f
based on the difference quotient method.
Theorem 3.15. (Interior H 2 -regularity) Let L be uniformly elliptic, with aij ∈ C 1 (Ω), bi
and c ∈ L∞ (Ω). Let f ∈ L2 (Ω). If u ∈ H 1 (Ω) is a weak solution of (3.33), then for any
Ω0 ⊂⊂ Ω we have u ∈ H 2 (Ω0 ), and
(3.34) kukH 2 (Ω0 ) ≤ C (kukL2 (Ω) + kf kL2 (Ω) )
where the constant C depends only on n, Ω0 , Ω and the coefficients of L.
Proof. Set q = f − ni=1 bi Di u − cu. Since u is a weak solution of (3.33), (by the similar
P
definition as above), this means that
Z X n Z
(3.35) aij Di uDj ϕdx = qϕdx ∀ ϕ ∈ H01 (Ω), supp ϕ ⊂⊂ Ω.
Ω i,j=1 Ω
Step 1: (Interior H 1 -estimate). Take any Ω00 ⊂⊂ Ω. Choose a cutoff function ζ ∈ C0∞ (Ω)
with 0 ≤ ζ ≤ 1 and ζ|Ω00 = 1. We take ϕ = ζ 2 u in (3.35) and perform elementary calculations
using the ellipticity condition, to discover
Z Z
ζ 2 |∇u|2 dx ≤ C (f 2 + u2 )dx.
Ω Ω
Thus
(3.36) kukH 1 (Ω00 ) ≤ C(kf kL2 (Ω) + kukL2 (Ω) ),
where the constant C depends on Ω00 .
3.5. Regularity 71
By using an induction argument, we can also get higher regularity for the solution.
Theorem 3.16. (Higher interior regularity) Let L be uniformly elliptic, with aij ∈
C k+1 (Ω), bi , c ∈ C k (Ω), and f ∈ H k (Ω). If u ∈ H 1 (Ω) is a weak solution of Lu = f , then
for any Ω0 ⊂⊂ Ω we have u ∈ H k+2 (Ω0 ) and
(3.38) kukH k+2 (Ω0 ) ≤ C (kukL2 (Ω) + kf kH k (Ω) )
where the constant C depends only on n, Ω0 , Ω and the coefficients of L.
72 3. Second-Order Linear Elliptic Equations
Proof. Suppose we have proved this theorem for k. Now assume aij ∈ C k+2 (Ω), bi , c ∈
C k+1 (Ω), f ∈ H k+1 (Ω) and u ∈ H 1 (Ω) is a weak solution of Lu = f. Then, by the induction
k+2 k+3
assumption, u ∈ Hloc (Ω), with the estimate (3.38). We want to show u ∈ Hloc (Ω). Fix
0 00
Ω ⊂⊂ Ω ⊂⊂ Ω and a multiindex α with |α| = k + 1. Let
ũ = Dα u ∈ H 1 (Ω00 ).
Given any ṽ ∈ C0∞ (Ω00 ), let ϕ = (−1)|α| Dα ṽ be put into the identity B1 [u, ϕ] = (f, ϕ)L2 (Ω)
and perform some elementary integration by parts, and eventually we discover
B1 [ũ, ṽ] = (f˜, ṽ)L2 (Ω) ,
where
n n
X α X α−β X
f˜ := Dα f − − (D aij Dβ uxi )xj + Dα−β bi Dβ uxi + Dα−β cDβ u .
β
β≤α,β6=α i,j=1 i=1
That is, ũ ∈ H 1 (Ω00 ) is a weak solution of Lũ = f˜ on Ω00 . (This is equivalent to differentiating
the equation Lu = f with Dα -operator.) We have f˜ ∈ L2 (Ω00 ), with, in light of the induction
assumption on the H k+2 (Ω00 )-estimate of u,
kf˜kL2 (Ω00 ) ≤ C(kf kH k+1 (Ω00 ) + kukH k+2 (Ω00 ) ) ≤ C(kf kH k+1 (Ω) + kukL2 (Ω) ).
Therefore, by Theorem 3.15, ũ ∈ H 2 (Ω0 ), with the estimate
kũkH 2 (Ω0 ) ≤ C(kf˜kL2 (Ω00 ) + kũkL2 (Ω00 ) ) ≤ C(kf kH k+1 (Ω) + kukL2 (Ω) ).
This exactly proves u ∈ H k+3 (Ω0 ) and the corresponding estimate (3.38) with k + 1.
3.5.2. Boundary Regularity. We now study the regularity up to the boundary. For this
purpose we need certain smoothness of the boundary. We have the following result.
Theorem 3.17. (Global H 2 -regularity) Assume in addition to the assumptions of The-
orem 3.15 that aij ∈ C 1 (Ω̄) and ∂Ω ∈ C 2 . If u ∈ H01 (Ω) is a weak solution to Lu = f , then
u ∈ H 2 (Ω), and
(3.39) kukH 2 (Ω) ≤ C(kukL2 (Ω) + kf kL2 (Ω) )
where the constant C depends only on n, kaij kW 1,∞ (Ω) , kbi kL∞ (Ω) , kckL∞ (Ω) and ∂Ω.
Since u = 0 along {xn = 0} and ζ = 0 near the curved portion of ∂Ω, we see ϕ ∈ H01 (Ω).
Therefore, we can use this ϕ as a test function in (3.35) as we did in the Step 2 above. The
end result is that
Dk u ∈ H 1 (Ω0 ) (k = 1, 2, · · · , n − 1)
with the estimate
X n
(3.40) kDkl ukL2 (Ω0 ) ≤ C(kf kL2 (Ω) + kukL2 (Ω) ).
k,l=1,k+l<2n
3. We must estimate kDnn ukL2 (Ω0 ) . We now use the fact that the assumption aij ∈ C 1
and the interior regularity imply the equation Lu = f is satisfied almost everywhere in
Ω; in this case we say u ∈ Hloc 2 (Ω) is a strong solution. Since the ellipticity implies
ann (x) ≥ θ > 0, we can actually solve Dnn u from the equation Lu = f in terms of Dij u
and Di u with i + j < 2n, i, j = 1, 2, · · · , n. This way we deduce the pointwise estimate
X n
|Dnn u| ≤ C |Dij u| + |∇u| + |u| + |f | .
i,j=1,i+j<2n
Hence, by (3.40),
kukH 2 (Ω0 ) ≤ C(kf kL2 (Ω) + kukL2 (Ω) ).
4. Now it is standard to treat smooth domains by locally flattening the boundary. Since
∂Ω is C 2 , at each point x0 ∈ ∂Ω, we have a small ball B(x0 , r) and a C 2 map y = Φ(x),
with Φ(x0 ) = 0, that maps B(x0 , r) bijectively onto a domain in the y space such that
Φ(Ω ∩ B(x0 , r)) ⊂ {y ∈ Rn | yn > 0}.
We assume the inverse of this map is x = Ψ(y). Both Ψ and Φ are C 2 . Choose s > 0 so small
that the half-ball V := B(0, s) ∩ {yn > 0} lies in Φ(Ω ∩ B(x0 , r)). Set V 0 = B(0, s/2) ∩ {yn >
0}. Finally define
v(y) = u(Ψ(y)) (y ∈ V ).
1
Then v ∈ H (V ) and v = 0 on ∂V ∩ {yn = 0} (in the sense of trace). Moreover, u(x) =
v(Φ(x)) and hence
n
X
uxi (x) = vyk (Φ(x))Φkxi (x) (i = 1, 2, · · · , n).
k=1
Z
= aij (Ψ(y))vyk (y)Φkxi (Ψ(y))ϕyl (y)Φlxj (Ψ(y))
V
+ bi (Ψ(y))vyk (y)Φkxi (Ψ(y))ϕ(y) + c(Ψ(y))v(y)ϕ(y) |I(y)|dy.
|I|y
Since ϕyl |I| = ζyl − |I|l ζ, we have
Z
(3.42) B1 [u, w] = ãij (y)vyk (y)ζyk (y) + b̃i (y)vyk (y)ζ(y) + c̃(y)v(y)ζ(y) dy := B̃[v, ζ],
V
and
n n
X X |I|yl
b̃k (y) = bi (x)Φkxi (x) − aij (Ψ(y))Φkxi (Ψ(y))Φlxj (Ψ(y)) (k = 1, 2, · · · , n).
|I|
i=1 i,j,l=1
6. We easily have that ãkl ∈ C 1 (V̄ ), b̃k , c̃ ∈ L∞ (V ). We now check that the operator
M is uniformly elliptic in V . Indeed, if y ∈ V and ξ ∈ Rn , then, again with x = Ψ(y),
n
X n X
X n n
X
ãkl (y)ξk ξl = ars (x)Φkxr Φlxs ξk ξl = ars (x)ηr (x)ηs (x) ≥ θ|η(x)|2 ,
k,l=1 r,s=1 k,l=1 r,s=1
That is, η(x) = ξDΦ(x). Hence ξ = η(x)DΨ(y) with y = Φ(x). So |ξ| ≤ C|η(x)| for some
constant C. This shows that
n
X
ãkl (y)ξk ξl ≥ θ|η(x)|2 ≥ θ0 |ξ|2
k,l=1
for some constant θ0 > 0 and all y ∈ V and ξ ∈ Rn . By the result with flat boundary in
Step 1, we have
kvkH 2 (V 0 ) ≤ C(kgkL2 (V ) + kvkL2 (V ) ).
Consequently, with O0 = Ψ(V 0 ), using the fact Φ, Ψ are of C 2 , we deduce
(3.43) kukH 2 (O0 ) ≤ C(kf kL2 (Ω) + kukL2 (Ω) ).
Note that x0 ∈ Ψ(B(0, s/2)) := G0 , which is an open set containing open set O0 .
3.5. Regularity 75
7. Since ∂Ω is compact, we can find finitely many open sets Oi0 ⊂ G0i (i = 1, 2, · · · , k)
such that ∂Ω ⊂ ∪ki=1 G0i . Then there exists a δ > 0 such that
k
[
F := {x ∈ Ω | dist(x, ∂Ω) ≤ δ} ⊂ Oi0 .
i=1
Then U = (Ω \ F ) ⊂⊂ Ω. By (3.43), we have
kukH 2 (F ) ≤ C(kf kL2 (Ω) + kukL2 (Ω) ).
By interior regularity,
kukH 2 (U ) ≤ C(kf kL2 (Ω) + kukL2 (Ω) ).
Combining these two estimates, we deduce (3.39).
Theorem 3.18. (Higher global regularity) Let L be uniformly elliptic, with aij ∈
C k+1 (Ω̄), bi , c ∈ C k (Ω̄), f ∈ H k (Ω), and ∂Ω ∈ C k+2 . Then a weak solution u of Lu = f
satisfying u ∈ H01 (Ω) belongs to H k+2 (Ω), and
(3.44) kukH k+2 (Ω) ≤ C(kukL2 (Ω) + kf kH k (Ω) )
where the constant C is independent of u and f . Furthermore, if the only weak solution
u ∈ H01 (Ω) of Lu = 0 is u ≡ 0, then, whenever Lu = f ,
(3.45) kukH k+2 (Ω) ≤ Ckf kH k (Ω)
where C is independent of u and f .
3. We need to extend (3.48) to all β with |β| = k + 3. Fix k, we prove (3.48) for all β
by induction on j = 0, 1, · · · , k + 2 with βn ≤ j. We have already shown it for j = 0, 1, 2.
Assume we have shown it for j. Now assume β with |β| = k + 3 and βn = j + 1. Let us
k+3
write β = γ + δ, for δ = (0, · · · , 0, 2) and so |γ| = k + 1. Since u ∈ Hloc (Ω) and Lu = f in
Ω, we have Dγ Lu = Dγ f a.e. in Ω. Now
Dγ Lu = ann Dβ u + R,
where R is the sum of terms involving at most j derivatives of u with respect to xn and
at most k + 3 derivatives in all. Since ann ≥ θ, we can solve Dβ u in terms of R and Dγ f ;
hence,
kDβ ukL2 (Ωt ) ≤ C(kf kH k+1 (Ω) + kukL2 (Ω) ).
By induction, we deduce (3.48), which proves
kukH k+3 (Ωt ) ≤ C(kf kH k+1 (Ω) + kukL2 (Ω) ).
This estimate in turn completes the induction process on k, begun in step 2. This proves
(3.46).
4. As above, we can cover the domain Ω̄ by finitely many small balls and use the method
of flattening the boundary to eventually deduce (3.44).
5. We prove the last statement (3.45) for k = 0; the more general case is similar. In
view of (3.44), it suffices to show that
kukL2 (Ω) ≤ Ckf kL2 (Ω)
If to the contrary this inequality is false, there would exist sequences un ∈ H 2 (Ω) ∩ H01 (Ω)
and fn ∈ L2 (Ω) for which kun k2 = 1 and kfn k2 → 0. By (3.44) we have kun k2,2 ≤ C. Thus
we can assume that un converges weakly to u in H 2 (Ω) and strongly in L2 (Ω). For fixed
v ∈ H01 (Ω), the functional l(u) = B[u, v] ∈ (H 2 (Ω))∗ , and so by passing to the limit in
Z
l(un ) = B[un , v] = fn vdx for all v ∈ H01 (Ω)
Ω
we see that B[u, v] = 0 for all v ∈ H01 (Ω) and thus u is a weak solution of Lu = 0. Hence,
u ≡ 0 by weak uniqueness. This contradicts kuk2 = limn→∞ kun k2 = 1.
where Aαβ ∞
ij ∈ L (Ω). Consider the linear partial differential system
−Dα Aαβ j
= g i − Dα fαi , i = 1, 2, · · · , N,
(3.49) ij (x) Dβ u
3.6. Regularity for Linear Systems* 77
(H2) Aαβ αβ i j 2 2
ij are constants, Aij pα pβ q q ≥ ν |p| |q| .
(H3) Aαβ αβ i j 2 2
ij ∈ C(Ω̄), Aij (x) pα pβ q q ≥ ν |p| |q| .
Under such a condition, we shall have the following Gårding inequality holds:
Z Z Z
2
(3.51) A(x)Dψ · Dψ ≥ ν0 |Dψ| − ν1 |ψ|2 , ∀ ψ ∈ H01 (BR ; RN ),
BR BR BR
where ν0 > 0, ν1 ≥ 0 are constants. For, under the hypothesis (H1) or (H2) the Gårding
inequality (3.51) holds with ν0 = ν, ν1 = 0, and under (H3) the inequality (3.51) also holds
(see Theorem 3.8).
(3.49). Almost all the estimates pertaining to regularity of u are derived using test functions
of the form φ = η (u−λ), where η is a cut-off function which belongs to W01,∞ (Ω0 ) for certain
Ω0 ⊂⊂ Ω. Let Bρ ⊂⊂ BR ⊂⊂ Ω be concentric balls with center a ∈ Ω. Let
1
if 0 ≤ t ≤ ρ,
R−t
θ(t) = R−ρ if ρ ≤ t ≤ R,
0 if t > R.
Let ζ = ζρ,R (x) = θ(|x − a|). Then ζ ∈ W01,∞ (Ω) with supp ζ ⊆ B̄R and
χρ,R
(3.52) 0 ≤ ζ ≤ 1, ζ|Bρ ≡ 1, |Dζ| ≤ ,
R−ρ
where χρ,R = χBR \Bρ (x) is the characteristic function of BR \ Bρ . Define
ψ = ζ (u − λ), φ = ζ 2 (u − λ) = ζ ψ.
Then ψ, φ ∈ H01 (BR ; RN ) and
Dφ = ζDψ + ψ ⊗ Dζ, Dψ = ζDu + (u − λ) ⊗ Dζ.
Using φ as a test function in (3.50) yields
Z Z
(g · φ + f · Dφ) = A(x)Du · Dφ
BR BR
Z Z
= A(x)Du · ζDψ + A(x)Du · ψ ⊗ Dζ.
BR BR
78 3. Second-Order Linear Elliptic Equations
Note that
A(x)Dψ · Dψ = A(x)Du · ζDψ + A(x)(u − λ) ⊗ Dζ · Dψ,
A(x)Du · ψ ⊗ Dζ = A(x)ζDu · (u − λ) ⊗ Dζ
= A(x)Dψ · (u − λ) ⊗ Dζ
− A(x)(u − λ) ⊗ Dζ · (u − λ) ⊗ Dζ.
Therefore, it follows that
Z Z
A(x)Dψ · Dψ = A(x)(u − λ) ⊗ Dζ · Dψ
BR BR
Z Z
+ (g · φ + f · Dφ) − A(x)Dψ · (u − λ) ⊗ Dζ
BR BR
Z
+ A(x)(u − λ) ⊗ Dζ · (u − λ) ⊗ Dζ.
BR
(Note that the first and third terms on the righthand side would cancel out if A(x)ξ · η is
symmetric in ξ, η.) Then, by (3.51), it follows that
Z Z Z
2
ν0 |Dψ| ≤ A(x)Dψ · Dψ + ν1 |ψ|2
BR B BR
ZR
χρ,R |u − λ|
Z Z
≤ g·φ + |f | · |Dψ| + |f | ·
BR BR BR R−ρ
χρ,R |u − λ| χρ,R |u − λ|2
Z Z Z
+ C |Dψ| · +C + ν1 |u − λ|2 .
BR R−ρ BR (R − ρ)2 BR
Using the Cauchy inequality with , we deduce
"Z #
|u − λ|2
Z Z Z Z
2 2 2
(3.53) |Dψ| ≤ C 2
+ g·φ + |f | + ν1 |u − λ| .
BR BR \Bρ (R − ρ) BR BR BR
Since ψ|Bρ = u − λ, this last estimate (3.53) proves the following theorems.
1 (Ω; RN ) be a weak solution of (3.49). Assume either condition
Theorem 3.20. Let u ∈ Hloc
(H1) or (H2) holds. Then
"Z #
Z
|u − λ| 2 Z Z
(3.54) |Du|2 ≤ C 2
+ g · ζ 2 (u − λ) + |f |2
Bρ BR \Bρ (R − ρ) BR BR
for all concentric balls Bρ ⊂⊂ BR ⊂⊂ Ω and constants λ ∈ RN , where ζ = ζρ,R and C > 0
is a constant depending on the L∞ -norm of Aαβ
ij .
see later that this term needs a special consideration when we deal with higher regularity
for weak solutions, especially when g is of the form of quotient difference.
We choose Z
1
λ= u(x) dx.
|BR \ BR/2 | BR \BR/2
Therefore Z Z
2
|Du| ≤ C |Du|2 .
BR/2 BR \BR/2
Adding C B |Du|2 to both sides of this inequality (a.k.a. the hole-filling technique of
R
R/2
Widman), we obtain
Z Z
2 C
|Du| ≤ |Du|2 .
BR/2 C + 1 BR
Letting R → ∞ we have
Z Z
C
|Du|2 dx ≤ |Du|2 dx.
Rn C +1 Rn
C
|Du|2 = 0 and thus Du ≡ 0; hence u ≡ constant.
R
Since C+1 < 1 we have Rn
Corollary 3.24. Assume either condition (H1) or (H2) holds. Then any bounded weak
1 (R2 ; RN ) to (3.57) for n = 2 must be constant.
solution u ∈ Hloc
Proof. Let |u| ≤ M ; then by the Caccioppoli inequality (3.56) with λ = 0 we have
Z
|Du|2 dx ≤ CM < ∞, ∀ R > 0.
BR/2
holds for all φ ∈ W01,2 (Ω0 ; RN ). If 0 < |h| < dist(Ω0 ; ∂Ω) we have
Z
A(x + hes ) Du(x + hes ) · Dφ(x) dx
Ω
Z Z
= g(x + hes ) · φ(x) dx + f (x + hes ) · Dφ(x) dx.
Ω Ω
Subtract two equations and divide by h to get
Z Z
h
A(x + hes ) DDs u · Dφ = Dsh g(x) · φ(x) dx
Ω
ZΩ Z
+ Dsh f (x) · Dφ(x) dx − Dsh A(x) Du(x) · Dφ(x) dx.
Ω Ω
Assume that Gårding’s inequality (3.51) holds. Then we can invoke the estimate (3.53)
with λ = 0, ρ = R/2 to obtain
Z hZ Z
2 1 h 2
(3.59) |Dψ| ≤ C 2
|Ds u| + Dsh g · φ
BR BR R BR
Z
i
+ |Dsh f |2 + ν1 |Dsh u|2 + |Dsh A|2 |Du|2 ,
BR
and thus Ds Du exists and belongs to L2 (BR/2 ; MN ×n ) for all s = 1, 2, · · · , n. This implies
2 (Ω; RN ). Therefore, we have proved the following theorem.
u ∈ Hloc
Theorem 3.25. Suppose A ∈ C(Ω̄) is Lipschitz continuous and the Gårding inequality
1 (Ω; MN ×n ) and u ∈ H 1 (Ω; RN ) is a weak solution
(3.51) holds. If g ∈ L2loc (Ω; RN ), f ∈ Hloc loc
of the system
− div(A(x) Du) = g − div f
2 (Ω; RN ).
then u ∈ Hloc
The following higher regularity result can be proved by the standard bootstrap method.
1 (Ω; RN ) is a weak solution of the system
Theorem 3.26. Suppose u ∈ Hloc
− div(A(x) Du) = g − div f
with A ∈ C k,1 (Ω̄) (that is, Dk A is Lipschitz continuous) satisfying the Gårding inequality
(3.51) and g ∈ Hlock (Ω; RN ), f ∈ H k+1 (Ω; MN ×n ). Then u ∈ H k+2 (Ω; RN ).
loc loc
Proof. Let ψ ∈ C0∞ (Ω; RN ); then we use φ = Ds ψ as a test function for the system to
obtain Z Z Z
Ds (A(x) Du) · Dψ dx = Ds g · ψ + Ds f · Dψ.
Ω Ω Ω
Since Ds (A(x) Du) = (Ds A) Du + A(x) DDs u we thus have
Z Z Z
A(x) D(Ds u) · Dψ = Ds g · ψ + Ds f − (Ds A) Du · Dψ.
Ω Ω Ω
Proof. By the Caccioppoli-type inequality, we have for any weak solution u of system (3.61)
Z Z
2 C
|Du| dx ≤ |u|2 dx.
Bρ (R − ρ)2 BR
k,2
The regularity result shows that u ∈ Wloc (Ω; RN ) for all k and then it follows that any
derivative Dk u is also a weak solution of (3.61). Therefore, the conclusion will follow from
a successive use of the above Caccioppoli inequality with a finite number of R/2 = ρ1 <
ρ2 < · · · < ρK = R.
In exactly the same way as the scalar equations, we obtain the global H 2 -regularity.
Theorem 3.28. (Global H 2 -regularity) Let ∂Ω be C 2 . Suppose A ∈ C 1 (Ω̄) and the
Gårding inequality (3.51) holds. If g ∈ L2 (Ω; RN ), f ∈ H 1 (Ω; MN ×n ) and u ∈ H 1 (Ω; RN )
is a weak solution of the system
− div(A(x) Du) = g − div f
then u ∈ H 2 (Ω; RN ) and we have
(3.62) kukH 2 (Ω) ≤ C(kgkL2 (Ω) + kf kH 1 (Ω) + kukL2 (Ω) ),
where the constant C depends only on Ω and the coefficients.
3.6.3. Morrey and Campanato Spaces. Let Ω ⊂ Rn be a bounded open domain. For
x ∈ Rn , ρ > 0 let
Ω(x, ρ) = {y ∈ Ω | |y − x| < ρ}.
Definition 3.15. For 1 ≤ p < ∞, λ ≥ 0 we define the Morrey space Lp,λ (Ω; RN ) by
Z
p,λ N p N −λ p
L (Ω; R ) = u ∈ L (Ω; R ) sup ρ |u(x)| dx < ∞ .
a∈Ω Ω(a,ρ)
0<ρ<diamΩ
We define a norm by
Z !1/p
−λ p
kukLp,λ (Ω;RN ) = sup ρ |u(x)| dx .
a∈Ω Ω(a,ρ)
0<ρ<diamΩ
so that kukLp,n ≤ C kuk∞ . Suppose now u ∈ Lp,n (Ω; RN ). Then by (3.63), (3.64)
Z
|u(a)| = lim − |u| ≤ C kukLp,n (Ω;RN ) .
ρ→0 Ω(a,ρ)
a ∈ Ω and 0 < ρ < ρ0 = min{1, diamΩ}. Suppose u ∈ Lq,µ (Ω; RN ). Then, by Hölder’s
inequality, for all a ∈ Ω, 0 < ρ < ρ0 < 1,
!p
Z Z q
p
1−
|u|p dx ≤ |Ω(a, ρ)| q |u|q dx
Ω(a,ρ) Ω(a,ρ)
n− np p
≤ Cρ q (kukqLq,µ (Ω;RN ) ρµ ) q
µp
+n− np
≤ Cρ q q kukpLq,µ (Ω;RN )
≤ C ρλ kukpLq,µ (Ω;RN ) ,
µp np
where we have used the assumption q +n− q ≥ λ and the fact 0 < ρ < 1. Therefore,
u ∈ Lp,λ (Ω; RN ).
Definition 3.16. For 1 ≤ p < ∞, λ ≥ 0 we define the Campanato space Lp,λ (Ω; RN ) by
Z
p,λ N p N −λ p
L (Ω; R ) = u ∈ L (Ω; R ) sup ρ |u − ua,ρ | dx < ∞ ,
a∈Ω Ω(a,ρ)
0<ρ<diamΩ
where ua,ρ is the average of u on Ω(a, ρ). Define the seminorm and norm by
Z !1/p
−λ p
[u]Lp,λ (Ω;RN ) = sup ρ |u − ua,ρ | dx ,
a∈Ω Ω(a,ρ)
0<ρ<diamΩ
kukLp,λ (Ω;RN ) = kukLp (Ω;RN ) + [u]Lp,λ (Ω;RN ) .
84 3. Second-Order Linear Elliptic Equations
Remark 3.17. For the estimates on linear elliptic systems of second-order partial differen-
tial equations, only spaces L2,λ (Ω; RN ) and L2,λ (Ω; RN ) are needed. The spaces with other
p ≥ 1 are useful for nonlinear problems; however, we shall not study the nonlinear problems
in this course.
Theorem 3.32. Both Lp,λ (Ω; RN ) and C 0,α (Ω̄; RN ) are Banach spaces.
Theorem 3.33. (a) For any p ≥ 1, λ ≥ 0, Lp,λ (Ω; RN ) ⊂ Lp,λ (Ω; RN ).
(b) For any 0 < α ≤ 1, C 0,α (Ω̄; RN ) ⊂ Lp,n+pα (Ω; RN ).
It turns out that we can exactly estimate the two terms on the right-hand side by
kvkLp (Ω(a,ρ)) ≤ ρλ/p kvkLp,λ (Ω;RN ) ,
Definition 3.18. We say that Ω ⊂ Rn is of type A if there exists a constant A > 0 such
that
(3.66) |Ω(a, ρ)| ≥ A ρn , ∀ a ∈ Ω, 0 < ρ < diamΩ.
This condition excludes that Ω may have sharp outward cusps; for instance, all Lipschitz
domains are of type A.
Lemma 3.34. Assume Ω is of type A and u ∈ Lp,λ (Ω; RN ). Then for any 0 < r < R <
∞, a ∈ Ω it follows that
− p1 λ
−n
|ua,R − ua,r | ≤ 2 A Rp r p · [u]Lp,λ (Ω;RN ) .
Proof.
1
|ua,R − ua,r | · |Ω(a, r)| p = kua,R − ua,r kLp (Ω(a,r))
≤ ku − ua,R kLp (Ω(a,r)) + ku − ua,r kLp (Ω(a,r))
≤ ku − ua,R kLp (Ω(a,R)) + ku − ua,r kLp (Ω(a,r))
λ λ
≤ [u]Lp,λ (Ω;RN ) R p + [u]Lp,λ (Ω;RN ) r p
λ
≤ 2 [u]Lp,λ (Ω;RN ) R p .
Hence the lemma follows from the assumption that |Ω(a, r)| ≥ A rn .
Theorem 3.35. If Ω is of type A then Lp,λ (Ω; RN ) ∼
= Lp,λ (Ω; RN ) for 0 ≤ λ < n.
Proof. We only need to show Lp,λ (Ω; RN ) ⊂ Lp,λ (Ω; RN ). Let u ∈ Lp,λ (Ω; RN ). Given any
a ∈ Ω, 0 < ρ < diamΩ, we have
kukLp (Ω(a,ρ)) ≤ ku − ua,ρ kLp (Ω(a,ρ)) + kua,ρ kLp (Ω(a,ρ))
λ n
≤ [u]Lp,λ (Ω;RN ) ρ p + C |ua,ρ | ρ p .
We choose an integer k large enough so that Ω(a, 2k ρ) = Ω. By Lemma 3.34, we have
k−1
X
|ua,ρ | ≤ |ua,2k ρ | + |ua,2j+1 ρ − ua,2j ρ |
j=0
k−1
− p1 λ
−n
X
≤ |uΩ | + 2A (2j+1 ρ) p (2j ρ) p · [u]Lp,λ (Ω;RN )
j=0
k
λ−n X
≤ |uΩ | + C ρ p [u]Lp,λ (Ω;RN ) · 2j(λ−n)/p
j=0
λ−n
≤ |uΩ | + C [u]Lp,λ (Ω;RN ) ρ p ,
where uΩ is the average of u on Ω and therefore |uΩ | ≤ C(Ω) kukLp (Ω;RN ) . Combining these
estimates, we deduce
λ n
kukLp (Ω(a,ρ)) ≤ C [u]Lp,λ (Ω;RN ) ρ p + C kukLp (Ω;RN ) ρ p
λ
and, by dividing both sides by ρ p and noting λ < n,
−λ
ρ p kukLp (Ω(a,ρ)) ≤ C [u]Lp,λ (Ω;RN ) + C(Ω) kukLp (Ω;RN ) .
86 3. Second-Order Linear Elliptic Equations
This proves
kukLp,λ (Ω;RN ) ≤ C(Ω) kukLp,λ (Ω;RN ) .
Remark 3.19. Note that Lp,n (Ω; RN ) ∼ 6 Lp,n (Ω; RN ) ∼
= = L∞ (Ω; RN ). For example, let
p = λ = 1, n = N = 1 and Ω = (0, 1). Then u = ln x is in L1,1 (0, 1) but not in L1,1 (0, 1) ∼
=
∞ p,n N ∼ N
L (0, 1). In fact, L (Ω; R ) = BM O(Ω; R ), which is called the John-Nirenberg space.
Theorem 3.36. (Campanato ’63) Let Ω be of type A. Then for n < λ ≤ n + p,
λ−n
Lp,λ (Ω; RN ) ∼
= C 0,α (Ω̄; RN ), α = ,
p
whereas for λ > n + p we have Lp,λ (Ω; RN ) = {constants}.
Proof. 1. Assume λ > n and v ∈ Lp,λ (Ω; RN ). For any x ∈ Ω and R > 0 we define
ṽ(x) = lim vx, R .
k→∞ 2k
We claim ṽ is well-defined and independent of R > 0. We first show the limit defining ṽ(x)
exists. We need to show the sequence {vx, R } is Cauchy. For h > k we have, by Lemma
2k
3.34,
h−1
X
|vx, R − vx, R | ≤ |vx, R − vx, R |
2h 2k 2j 2j+1
j=k
h−1
λ−n j(n−λ)
− p1
X
≤ 2A [v]Lp,λ (Ω;RN ) R p · 2 p ,
j=k
which, since λ > n, tends to zero if k, h → ∞. Therefore {vx, R } is Cauchy and the limit
2k
ṽ(x) exists. Also in the inequality above, if k = 0 and h → ∞ we also deduce
λ−n
(3.67) |vx,R − ṽ(x)| ≤ C [v]Lp,λ (Ω;RN ) · R p .
We now prove ṽ(x) is independent of R > 0. This follows easily since by Lemma 3.34
lim |vx, R − vx, r | = 0.
k→∞ 2k 2k
2. By Lebesgue’s differentiation theorem, we also have ṽ(x) = v(x) for almost every
x ∈ Ω. Therefore ṽ = v in Lp,λ (Ω; RN ). We claim ṽ ∈ C 0,α (Ω̄; RN ), where α = λ−n
p . To
show this, let x, y ∈ Ω and x 6= y. Let R = |x − y|. By (3.67) it follows that
|ṽ(x) − ṽ(y)| ≤ |ṽ(x) − vx,2R | + |ṽ(y) − vy,2R | + |vx,2R − vy,2R |
≤ C [v]Lp,λ (Ω;RN ) · Rα + |vx,2R − vy,2R |.
We need to estimate |vx,2R −vy,2R |. To this end, let S = Ω(x, 2R)∩Ω(y, 2R). Then Ω(x, R) ⊂
S and hence
|S| ≥ |Ω(x, R)| ≥ A Rn .
On the other hand, we have
1
|S| p · |vx,2R − vy,2R | = kvx,2R − vy,2R kLp (S)
≤ kvx,2R − vkLp (S) + kvy,2R − vkLp (S)
≤ kvx,2R − vkLp (Ω(x,2R)) + kvy,2R − vkLp (Ω(y,2R))
≤ 2 [v]Lp,λ (Ω;RN ) · (2R)λ/p .
3.6. Regularity for Linear Systems* 87
and hence
|ṽ(x) − ṽ(y)| ≤ C [v]Lp,λ (Ω;RN ) · |x − y|α .
This shows
[ṽ]C 0,α (Ω̄;RN ) ≤ C [v]Lp,λ (Ω;RN ) .
Finally, observe that, by (3.67) with R = diamΩ,
3. We have thus proved that if λ > n then every v ∈ Lp,λ (Ω; RN ) has a representation
ṽ which belongs to C 0,α (Ω̄; RN ) with α = (λ − n)/p. If λ > n + p then α > 1 and any u ∈
C 0,α (Ω̄; RN ) must be a constant (why?). The proof of Campanato’s theorem is complete.
In order to use the Campanato spaces for elliptic systems, we also need some local
version of these spaces. To disperse some technicalities, we prove the following lemma.
nondecreasing in subsets E ⊂⊂ Ω.
ZE E
= |u − uF |2 − |E| · |uF − uE |2
E
Z
≤ |u − uF |2 .
F
holds for all balls Bρ ⊂⊂ Ω. Then for any subdomain Ω0 ⊂⊂ Ω we have u ∈ Lp,λ (Ω0 ; RN )
and moreover
kukLp,λ (Ω0 ;RN ) ≤ C(Ω0 ) [Cu1/p + kukLp (Ω0 ;RN ) ].
Proof. Let Ω0 ⊂⊂ Ω be given. We will show u ∈ Lp,λ (Ω0 ; RN ). Let d = dist(Ω0 ; ∂Ω). Given
any a ∈ Ω0 and 0 < ρ < diam(Ω0 ). If ρ < dist(a; ∂Ω) we have by the previous lemma,
Z Z
p
|u − uΩ0 (a,ρ) | dx ≤ |u − uBρ (a) |p dx ≤ Cu ρλ .
Ω0 (a,ρ) Bρ (a)
0, βp
Then for any Ω0 ⊂⊂ Ω of type A, we have u ∈ C (Ω̄0 ; RN ).
3.6. Regularity for Linear Systems* 89
Then there exist positive constants σ0 and C depending only on n such that
Z σ
(3.71) exp |u − uΩ | dx ≤ C (diamΩ)n ,
Ω K
where σ = σ0 |Ω| (diamΩ)−n .
Remark 3.20. The set of all functions u ∈ W 1,1 (Ω; RN ) satisfying (3.70) is the space
BM O(Ω; RN ) introduced by John and Nirenberg, and for Ω cubes or balls it follows that
BM O(Ω; RN ) ∼
= Lp,n (Ω; RN ), ∀ p ≥ 1.
For the proof of all these results and more on BM O-spaces, we refer to Gilbarg-Trudinger’s
book for a proof based on the Riesz potential, and Giaquinta’s book on the Calderon-
Zygmund cube decomposition.
3.6.4. Estimates for systems with constant coefficients. We consider systems with
constant coefficients. Let A = Aαβ
ij be constants satisfying hypothesis (H2). We first have
some Campanato estimates for homogeneous systems.
1 (Ω; RN ) be a weak solution of
Theorem 3.41. Let u ∈ Hloc
(3.72) Dα (Aαβ j
ij Dβ u ) = 0, i = 1, 2, ..., N.
where y ∈ D ≡ {y ∈ Rn | a + Ry ∈ Ω}, which includes B̄1 (0) in the y-space. Note that
1 (D; RN ) is a weak solution of
v ∈ Hloc
Dyα (Aαβ j
ij Dyβ v ) = 0.
where we have chosen integer k > n/2 and used the Sobolev embedding W k,2 (B1/2 (0); RN ) ,→
C 0,α (B1/2 (0); RN ) for some 0 < α < 1. Now if t ≥ 1/2 we easily have
Z Z
2 n n
|v| dy ≤ 2 t |v|2 dy.
Bt (0) B1 (0)
this proves (3.73). Note that Du is also a weak solution of (3.72); therefore, by (3.73) it
follows that
Z Z
2 ρ n
|Du| dx ≤ C(n) ( ) · |Du|2 dx, ∀ ρ < R < dist(a; ∂Ω).
Bρ (a) R BR (a)
Suppose 0 < ρ < R/2. Then we use the Poincaré inequality, the previous estimate and the
Caccioppoli inequality to obtain
Z Z
2 2
|u − uBρ | dx ≤ c(n) ρ · |Du|2 dx
Bρ Bρ
Z
ρ
≤ C(n) ρ2 ( )n · |Du|2 dx
R BR/2
Z
ρ
≤ C(n) ( )n+2 · |u − uBR |2 dx.
R BR
3.6. Regularity for Linear Systems* 91
In (3.73) and (3.74), if we let R → dist(a; ∂Ω), we see that both estimates also hold for
all balls Bρ ⊂⊂ BR ⊂ Ω. We state this fact as follows.
Corollary 3.42. Both estimates (3.73) and (3.74) hold for all balls Bρ ⊂⊂ BR ⊂ Ω.
In the following, we consider the nonhomogeneous elliptic systems with constant coef-
ficients:
(3.75) Dα (Aαβ j i
ij Dβ u ) = Dα fα , i = 1, 2, · · · .
and hence
Z Z
|Du − Dv|2 dx ≤ Cν |f − fBR |2 dx ≤ Cν [f ]2L2,λ (Ω00 ;MN ×n ) · Rλ .
BR BR
Let Z
Φ(ρ) = |Du − (Du)Bρ |2 dx.
Bρ
Therefore
[Du]L2,λ (Ω0 ;MN ×n ) ≤ C kDukL2 (Ω00 ;MN ×n ) + [f ]L2,λ (Ω00 ;MN ×n ) .
The proof is complete.
3.6.5. Schauder estimates for systems with variable coefficients. We now study
the local regularity of weak solutions of systems with variable coefficients. We first prove
the regularity in the Morrey space L2,λ
loc (Ω) for the gradient of the weak solutions.
Suppose f ∈ L2,λ
loc (Ω; M
N ×n ) and 0 ≤ λ < n. Then Du ∈ L2,λ (Ω; MN ×n ).
loc
Proof. Let Ω0 ⊂⊂ Ω00 ⊂⊂ Ω. Let a ∈ Ω0 and BR (a) = BR ⊂ Ω00 . Using the standard Korn’s
freezing coefficients device, u is a weak solution of system with constant coefficients
div(A(a) Du) = div F, F = f + (A(a) − A(x)) Du.
Let v ∈ H 1 (BR ; RN ) be the solution of the Dirichlet problem
(
div(A(a) Dv) = 0 in BR ,
v|∂BR = u.
Then, as before, using (3.73) instead of (3.74) we have
Z Z Z
2 ρ n 2
|Du| ≤ c · ( ) |Du| + C |D(u − v)|2
Bρ R BR BR
Z Z
ρ n 2
≤ c·( ) |Du| + C |F |2
R BR BR
Z Z Z
ρ
≤ c · ( )n |Du|2 + C |f |2 + C ω(R) |Du|2
R BR BR BR
h ρ iZ
≤ c ( )n + ω(R) |Du|2 + C kf k2L2,λ (Ω00 ;MN ×n ) Rλ ,
R BR
94 3. Second-Order Linear Elliptic Equations
We choose R0 > 0 sufficiently small so that ω(R) < 0 for all R < R0 , where 0 is the
constant appearing in the Campanato lemma above. Therefore,
Z
|Du|2 dx ≤ C(Ω0 , Ω00 ) kDuk2L2 (Ω00 ;MN ×n ) + kf k2L2,λ (Ω00 ;MN ×n ) ρλ .
Bρ
We now study the regularity of the gradient of weak solutions in the Hölder spaces.
2,n+2µ
This is done by proving the regularity of gradient in the Campanato space Lloc (Ω) for
some µ ∈ (0, 1).
(3.79) Dα (Aαβ j i
ij (x) Dβ u ) = Dα fα .
0,µ 0,µ
Suppose f ∈ Cloc (Ω; MN ×n ). Then Du ∈ Cloc (Ω; MN ×n ).
Z Z
2 ρ n+2
|Du − (Du)Bρ | ≤ A ( ) |Du − (Du)BR |2 + B Rn+2µ− .
Bρ R BR
Finally we remark that the following higher order regularity result can be easily deduced.
Theorem 3.48. Let k ≥ 0, 0 < µ < 1 and Aαβ ij ∈ C
k,µ (Ω) satisfy the hypothesis (H3) and
1,2
u ∈ Wloc (Ω; RN ) be a weak solution of system
(3.80) Dα (Aαβ j i
ij (x) Dβ u ) = Dα fα .
k,µ k+1,µ
Suppose f ∈ Cloc (Ω; MN ×n ). Then u ∈ Cloc (Ω; RN ).
We now consider the regularity up to the boundary. In what follows, we assume the
boundary ∂Ω of the domain Ω is of C 1,µ ; that is, for any x0 ∈ ∂Ω, there exist an open set
U ⊂ Rn containing x0 and a C 1,µ -diffeomorphism y = G : U → Rn such that
G(x0 ) = 0, G(U ∩ Ω) = B1+ = {y ∈ Rn | |y| < 1, yn > 0};
Let A(x) satisfy the condition (H3). If u ∈ H 1 (Ω; RN ) is a weak solution to the problem
div(A(x) Du) = div f, u|∂Ω = g,
then u ∈ C 1,µ (Ω̄; RN ).
96 3. Second-Order Linear Elliptic Equations
Linear Evolution
Equations
This chapter studies various linear partial differential equations that involve time. We call
these equations linear evolution equations. We will study two major types of evolution
equations of second-order: parabolic and hyperbolic equations. Two methods will be used:
Galerkin method and Semigroup method.
97
98 4. Linear Evolution Equations
∂
Definition 4.1. We say the operator ∂t + L is (uniformly) parabolic on ΩT if there
exists a constant θ > 0 such that
n
X
(4.4) aij (x, t)ξi ξj ≥ θ|ξ|2 for all (x, t) ∈ ΩT and ξ ∈ Rn .
i,j=1
Note that for each fixed time t ∈ [0, T ] the operator Lu is uniformly elliptic in x ∈ Ω.
4.1.1. Weak Solutions. We consider the case that Lu has the divergence form (4.2). We
assume
aij , bi , c ∈ L∞ (ΩT ) (i, j = 1, 2, · · · , n), f ∈ L2 (ΩT ), g ∈ L2 (Ω).
We will also assume aij = aji for i, j = 1, 2, · · · , n.
Introduce the time-dependent bilinear form
n
Z X n
X
(4.5) B[u, v; t] = aij (x, t)Di uDj v + bi (x, t)Di uv + c(x, t)uv dx
Ω i,j=1 i=1
Definition 4.2. A weak solution to Problem (4.1) is a function u ∈ L2 (0, T ; H01 (Ω)) with
weak time-derivative u0 ∈ L2 (0, T ; H −1 (Ω)) such that
(i) hu0 (t), vi + B[u(t), v; t] = (f (t), v) for each v ∈ H01 (Ω) and a.e. time t ∈ [0, T ], and
(ii) u(0) = g. (Note that u ∈ C([0, T ]; L2 (Ω)) and thus u(0) is well-defined in L2 (Ω).)
pairing ( , ) is the inner product in L2 (Ω). Note that the PDE can be written as
where the P
ut = g + ni=1 gxi i , with
0
n
X n
X
g0 = f − bi Di u − cu, gi = aij Di u (i = 1, 2, · · · , n).
i=1 i=1
(4.7) kut kH −1 (Ω) ≤ kGkL2 (Ω) ≤ C(kukH 1 (Ω) + kf kL2 (Ω) ) a.e. t ∈ [0, T ].
This estimate suggests it is reasonable to look for weak solutions u ∈ L2 (0, T ; H01 (Ω))
with weak time-derivative u0 (t) ∈ H −1 (Ω) for a.e. t ∈ [0, T ], which, by (4.7), also satisfies
u0 ∈ L2 (0, T ; H −1 (Ω)). In this case the first term in (4.6) should be reexpressed as hu0 (t), vi,
as the pairing of H −1 (Ω) and H01 (Ω).
4.1. Second-order Parabolic Equations 99
where αij (t) = B[wj , wi ; t] (i, j = 1, 2, · · · , k). Hence condition (4.10) becomes the initial
value problem for the ODE system on d(t) = (d1 (t), · · · , dk (t)):
k
αij (t)dj (t) = fi (t) ≡ (f (t), wi ),
X
d0i (t) + di (0) = (g, wi ) (i = 1, 2, · · · , k).
j=1
Note that the coefficients αij belong to L∞ (0, T ) and fi ∈ L2 (0, T ). The existence of a
unique solution d ∈ H 1 (0, T ) ⊂ C([0, T ]) is guaranteed by the (not so) standard existence
theory for ODE (think of approximating αij and fi by smooth functions first and then pass
to limits).
Proof. 1. From the uniform parabolicity, as in the elliptic case, there exist constants
β > 0, γ ≥ 0 such that
(4.12) βkvk2H 1 (Ω) ≤ B[v, v; t] + γkvk2L2 (Ω) ∀ v ∈ H01 (Ω), a.e. t ∈ [0, T ].
0
100 4. Linear Evolution Equations
Since η(0) = kuk (0)k2L2 (Ω) ≤ kgk2L2 (Ω) , we obtain the estimate
4. Finally we need to estimate ku0k kL2 (0,T ;H −1 (Ω)) . So, fix any v ∈ H01 (Ω), with kvkH01 (Ω) ≤
1. We write v = v 1 + v 2 , where v 1 ∈ Vk , and (v 2 , wi ) = 0 for all i = 1, 2, · · · , k. (That is, v 2
is in the L2 orthogonal complement of Vk .) Since {wi } are orthogonal in H01 (Ω), we have
kv 1 kH01 (Ω) ≤ kvkH01 (Ω) ≤ 1.
which, combining with the estimate in Step 3, derives the desired estimate.
4.1. Second-order Parabolic Equations 101
Proof. (Existence.) 1. According to the energy estimate (4.11), we see that {uk } is
bounded in L2 (0, T ; H01 (Ω)) and {u0k } is bounded in L2 (0, T ; H −1 (Ω)). Consequently there
exists a subsequence {ukm } of {uk } with km → ∞ and functions u ∈ L2 (0, T ; H01 (Ω)),
w ∈ L2 (0, T ; H −1 (Ω)) such that
(
ukm * u in L2 (0, T ; H01 (Ω)),
(4.15)
u0km * w in L2 (0, T ; H −1 (Ω)).
This equality then holds for all functions ψ ∈ L2 (0, T ; H01 (Ω)), as functions ζ of the given
form are dense in this space. We then take ψ(t) = ζ(t)v with ζ ∈ L2 (0, T ) and v ∈ H01 (Ω)
in (4.17) to obtain
Z T Z T
0
ζ(t) hu (t), vi + B[u(t), v; t] dt = ζ(t)(f (t), v)dt.
0 0
3. We need to show the initial data u(0) = g. In (4.16), (4.17), take ψ(t) = ζ(t)v with
ζ ∈ C 1 [0, T ], ζ(T ) = 0, ζ(0) = −1 and v ∈ H01 (Ω). Note that, since ψ 0 (t) = ζ 0 (t)v, we have
Z T Z T
0
huk (t), ψ(t)idt = (uk (0), v) − ζ 0 (t)(uk (t), v) dt,
0 0
Hence, in (4.16), let k = km → ∞, we eventually obtain (u(0), v) = (g, v), for all v ∈ H01 (Ω);
hence u(0) = g.
102 4. Linear Evolution Equations
4.1.5. Regularity. We now discuss the regularity of weak solutions when the initial data
and coefficients are more regular. Our eventual goal is to prove that the weak solution is
smooth, as long as the coefficients and initial data and the domain are all smooth. This
mirrors the regularity of elliptic equations.
Before we proceed, we prove the following useful result which is Problem 9 of Chapter
7 in Evans’s book; the proof follows a paper by Brezis and Evans (Arch. Rational Mech.
Analysis 71 (1979), 1–13).
Lemma 4.4. If Lu is uniformly elliptic with smooth coefficients, then there exist constants
β > 0, γ ≥ 0 such that
(4.19) βkuk2H 2 (Ω) ≤ (Lu, −∆u) + γkuk2L2 (Ω) ∀ u ∈ H01 (Ω) ∩ H 2 (Ω).
Proof. 1. Given a function u ∈ H 2 (Ω) ∩ H01 (Ω), we claim that there exists a sequence of
functions um in C 3 (Ω̄) vanishing on ∂Ω such that kum − ukH 2 (Ω) → 0; therefore, we only
need to prove (4.19) for functions u ∈ C 3 (Ω̄) vanishing on ∂Ω. To prove this claim, let
f = ∆u ∈ L2 (Ω) and choose fm ∈ C ∞ (Ω̄) so that kfm − f kL2 (Ω) → 0. Let um ∈ H01 (Ω)
be the weak solution of ∆um = fm in Ω. Then the global regularity theorem shows that
um ∈ H k (Ω) if ∂Ω is of C k and k ≥ 2. By the general Sobolev inequalities we know that if
n
k > n2 then um ∈ C k−[ 2 ]−1,γ (Ω̄) for some 0 < γ < 1. Hence if ∂Ω is of C k with k = 4 + [ n2 ],
then um ∈ C 3 (Ω̄). Clearly um = 0 on ∂Ω. Finally, since ∆(um − u) = fm − f , by the global
H 2 -estimate, we have
kum − ukH 2 (Ω) ≤ Ckfm − f kL2 (Ω) → 0.
The coefficients ck (y) are bounded (in fact, c2 (y) ≡ 0) and the coefficients akij (y), bki (y) are
smooth functions on D ∩ {yn ≥ 0} and satisfy for a constant θ > 0, with k = 1, 2,
akij (y) = akji (y),
akij (y)ξi ξj ≥ θ|ξ|2 (y ∈ D ∩ {yn ≥ 0}, ξ ∈ Rn ).
From this, an easy algebra proof shows that for all symmetric matrices C ∈ Mn×n
n
X n
X
(4.20) a1ij (y)cik a2kl (y)cjl ≥ θ2 c2ij (y ∈ D ∩ {yn ≥ 0}).
i,j,k,l=1 i,j=1
ψ(y)(L1 v, L2 v)dy.
R
We need to estimate D∩{yn >0}
3. We write the integrand as
ψ(y)(L1 v, L2 v) = ψa1ij vyi yj a2kl vyk yl + ψR,
where R is the term of the form
X X X
R= Aijk vyi yj vyk + Bij vyi vyj + Cij vyi yj v + c1 c2 v 2 ,
with bounded coefficients A, B, C. The leading term can be written as
ψa1ij vyi yj a2kl vyk yl = ψa1ij vyi yk a2kl vyj yl
+ (ψa1ij a2kl )yj vyi yk vyl − (ψa1ij a2kl )yk vyi yj vyl
+ (ψa1ij a2kl vyi yj vyl )yk − (ψa1ij a2kl vyi yk vyl )yj .
Therefore, by (4.20), we have
Z n Z
X
1 2 2
ψ(y)(L v, L v)dy ≥ θ ψ(y)vy2i yj dy
D∩{yn >0} i,j=1 D∩{yn >0}
n
X Z
(4.21) + ψa1ij a2kl (vyi yj vyl νk − vyi yk vyl νj ) dS
i,j,k,l=1 ∂(D∩{y n >0})
X n Z Z
− ε vy2i yj dy − Cε (|∇v|2 + v 2 ) dy,
i,j=1 D∩{yn >0} D∩{yn >0}
Z n−1
XZ
= ψa1nn a2nn vyn yn vyn dy 0 + 2 ψa1in a2nn vyi yn vyn dy 0 ,
D∩{yn =0} i=1 D∩{yn =0}
104 4. Linear Evolution Equations
and
n
X Z n Z
X
0
ψa1in a2kl vyi yk vyl dy = ψa1in a2kn vyi yk vyn dy 0
i,k,l=1 D∩{yn =0} i,k=1 D∩{yn =0}
Z n−1
XZ
0
= ψa1nn a2nn vyn yn vyn dy + ψa1in a2nn vyi yn vyn dy 0
D∩{yn =0} i=1 D∩{yn =0}
n−1
XZ
+ ψa1nn a2kn vyk yn vyn dy 0 .
k=1 D∩{yn =0}
Therefore
n
X Z
ψa1ij a2kl (vyi yj vyl νk − vyi yk vyl νj ) dS
i,j,k,l=1 ∂(D∩{yn >0})
n−1
XZ
= ψ(a1nn a2in − ψa1in a2nn )vyi yn vyn dy 0
i=1 D∩{yn =0}
n−1 Z
1X
= ψ(a1nn a2in − ψa1in a2nn )(vy2n )yi dy 0
2 D∩{yn =0}
i=1
n−1 Z
1X
=− [ψ(a1nn a2in − ψa1in a2nn )]yi vy2n dy 0 .
2 D∩{yn =0}
i=1
Hence
n
X Z
ψa1ij a2kl (vyi yj vyl νk − vyi yk vyl νj ) dS
i,j,k,l=1 ∂(D∩{yn >0})
n−1 Z Z
1X
= [ψ(a1nn a2in − ψa1in a2nn )]yi vy2n dy 0 ≤ C |∇v|2 dS
2 D∩{yn =0} ∂(D∩{yn >0})
i=1
n Z
X Z
≤ε vy2i yj dy + Cε |∇v|2 dy,
i,j=1 D∩{yn >0} D∩{yn >0}
where Cε depends on ζ. Clearly this estimate also holds when ζ ∈ Cc∞ (Ω).
7. We now cover ∂Ω by finitely many balls Bk := B(xk , rk ) with xk ∈ ∂Ω, k =
1, 2, · · · , N , that the local flattening of the boundary works. Let BN +1 := Ω\∪N k
k=1 B̄(x , rk /2).
PN +1
We find a partition of unity k=1 ζk = 1 subordinate to {B1 , B2 , · · · , BN +1 } with 0 ≤ ζk ≤
1 and supp ζk ⊂⊂ Bk . Then using (4.22) and a choice of small ε > 0 we deduce that
θ2
Z Z
(4.23) (Lu, −∆u) dx ≥ kD ukL2 (Ω) − C (|∇u|2 + u2 ) dx.
2 2
Ω 2 Ω
From this, the estimate (4.19) follows since kD2 ukL2 (Ω) is an equivalent norm for H 2 (Ω) ∩
H01 (Ω) in H 2 (Ω) and
k∇ukL2 (Ω) ≤ εkD2 ukL2 (Ω) + Cε kukL2 (Ω) ∀ u ∈ H 2 (Ω) ∩ H01 (Ω),
which can be seen from one of the homework problem.
Remark 4.4. Clearly, from the proof, the estimate (4.19) holds if we replace −∆u by
another uniformly elliptic operator M u.
Proof. Let {uk } be the Galerkin approximations satisfying (4.10) constructed as above
with {wi } being the complete collection of eigenfunctions with eigenvalues {λi } for −∆ on
H01 (Ω). As before, we assume {wi } is orthogonal on H01 (Ω) and orthonormal on L2 (Ω).
By the uniqueness theorem, the weak solution u is obtained as the limit of the Galerkin
approximations {uk }. We prove the theorem by deriving the same estimates for the ap-
proximate solutions uk independent of k.
106 4. Linear Evolution Equations
1. We first claim the following estimate for {uk }: for each t ∈ [0, T ],
(4.26) kuk (t)k2H 2 (Ω) ≤ C(kf (t)k2L2 (Ω) + ku0k (t)k2L2 (Ω) + kuk (t)k2L2 (Ω) ),
where C is a constant independent of k. To prove (4.26), note
(Luk (t), wi ) = B[uk (t), wi ; t] = (f (t) − u0k (t), wi ) (i = 1, 2, · · · , k).
Multiply this equation by λi di (t) and sum over i = 1, 2, · · · , k to deduce
(4.27) (Luk (t), −∆uk (t)) = B[uk (t), −∆uk (t); t] = (f (t) − u0k (t), −∆uk (t)),
since −∆uk (t) ∈ H01 (Ω). Then (4.26) follows from (4.27) and Lemma 4.4.
2. We multiply the equation in (4.10) by d0i (t) and sum i = 1, 2, · · · , k, to discover
(4.28) (u0k (t), u0k (t)) + B[uk (t), u0k (t); t] = (f (t), u0k (t)).
Write B[u, v; t] = A[u, v; t] + C[u, v; t], where
Z Xn
A[u, v; t] = aij (x, t)Di uDj v dx,
Ω i,j=1
(4.29)
n
Z !
X
C[u, v; t] = bi (x, t)(Di u)v + c(x, t)uv dx.
Ω i=1
Note that A[u, v; t] is a symmetric bilinear form on H01 (Ω) and for any functions u ∈
C 1 ([0, T ]; H01 (Ω)),
0 0 1 d
(4.30) A[u (t), u(t); t] = A[u(t), u (t); t] = A[u(t), u(t); t] − Ã[u(t), u(t); t] ,
2 dt
where
Z n
X
Ã[u, v; t] = a0ij (x, t)Di uDj v dx.
Ω i,j=1
where we have used (4.11) and the estimate A[uk (0), uk (0); 0] ≤ Ckuk (0)k2H 1 (Ω) ≤ Ckgk2H 1 (Ω) ,
0 0
since
k
X k
X
kuk (0)k2H 1 (Ω) = d2i (0)kwi k2H 1 (Ω) = (g, wi )2 kwi k2H 1 (Ω)
0 0 0
i=1 i=1
X∞
≤ (g, wi )2 kwi k2H 1 (Ω) = kgk2H 1 (Ω) .
0 0
i=1
ku0k k2L2 (0,T ;L2 (Ω)) + kuk k2L∞ (0,T ;H 1 (Ω)) ≤ C(kgk2H 1 (Ω) + kf k2L2 (ΩT ) ).
0 0
(ũ0k (t), ũk (t)) + B[ũk (t), ũk (t); t] = (f 0 (t), ũk (t)) − B̃[uk (t), ũk (t); t].
So,
1d
(4.35) kũk (t)k2L2 (Ω) + B[ũk (t), ũk (t); t] = (f 0 (t), ũk (t)) − B̃[uk (t), ũk (t); t].
2 dt
Notice that, for u ∈ H01 (Ω) ∩ H 2 (Ω), v ∈ H01 (Ω), integration by parts yields
Z n
X n
X
B̃[u, v; t] = v (−Dj a0ij Di u − a0ij Dij u) + b0i (x, t)Di u 0
+ c (x, t)u dx,
Ω i,j=1 i=1
and, hence, for a.e. t ∈ [0, T ] and all u ∈ H 2 (Ω), v ∈ H01 (Ω),
5. It remains to show u00 ∈ L2 (0, T ; H −1 (Ω))). To do so, take v ∈ H01 (Ω) with kvkH01 (Ω) ≤
1 and set v = v 1 + v 2 , as above with v 1 ∈ Vk and (v 2 , wi ) = 0 for i = 1, 2, · · · , k. Then, for
a.e. t ∈ [0, T ], by (4.34),
hu00k (t), vi = (u00k (t), v) = (u00k (t), v 1 ) = (f 0 (t), v 1 ) − B[u0k (t), v 1 ; t] − B̃[uk (t), v 1 ; t].
Hence, since kv 1 kH01 (Ω) ≤ 1,
|hu00k (t), vi| ≤ C(kf 0 (t)kL2 (Ω) + ku0k (t)kH01 (Ω) + kuk (t)kH01 (Ω) ).
This proves
ku00k (t)kH −1 (Ω) ≤ C(kf 0 (t)kL2 (Ω) + ku0k (t)kH01 (Ω) + kuk (t)kH01 (Ω) ).
So, squaring, integrating over t ∈ [0, T ] and using the estimates obtained above, we have
ku00k k2L2 (0,T ;H −1 (Ω)) ≤ C(kf 0 k2L2 (0,T ;L2 (Ω)) + ku0k k2L2 (0,T ;H 1 (Ω)) + kuk k2L2 (0,T ;H 1 (Ω)) )
0 0
We now study the higher regularity. For simplicity, we assume the coefficients of L
are smooth and independent of time t; we also assume the uniform parabolicity and the
smoothness of the domain Ω.
Theorem 4.6. (Higher regularity) Assume m ≥ 0 in an integer and
dk f
g ∈ H 2m+1 (Ω), ∈ L2 (0, T ; H 2m−2k (Ω)) (k = 0, 1, · · · , m).
dtk
Suppose the following m-th-order compatibility conditions hold:
g0 := g ∈ H01 (Ω), g1 := f (0) − Lg0 ∈ H01 (Ω),
(4.37) dm−1 f
· · · , gm := (0) − Lgm−1 ∈ H01 (Ω).
dtm−1
Then the weak solution u to (4.1) satisfies
dk u
∈ L2 (0, T ; H 2m−2k+2 (Ω)) (k = 0, 1, 2, · · · , m + 1),
dtk
with the estimate
m+1 m
!
X dk u X dk f
(4.38) k k kL2 (0,T ;H 2m−2k+2 (Ω)) ≤ C k k kL2 (0,T ;H 2m−2k (Ω)) + kgkH 2m+1 (Ω) .
dt dt
k=0 k=0
Proof. The proof is an induction on m, the case m = 0 being the conclusion (i) of Theorem
4.5 above. Assume now the theorem is valid for some integer m ≥ 0, and suppose then
dk f
g ∈ H 2m+3 (Ω), ∈ L2 (0, T ; H 2m+2−2k (Ω)) (k = 0, 1, · · · , m + 1)
dtk
and the (m + 1)-th-order compatibility conditions hold. Let ũ = u0 . Then the previous
theorem implies that ũ ∈ L2 (0, T ; H01 (Ω)) and ũ0 ∈ L2 (0, T ; H −1 (Ω)). Also ũ is the weak
solution to
˜
ũt + Lũ = f in ΩT ,
ũ = 0 on ∂Ω × [0, T ],
ũ = g̃ on Ω × {t = 0},
4.1.6. Maximum Principle. We now study some properties for classical (smooth) so-
lutions of parabolic equations. We include such a study here in order to compare with
the properties for classical solutions of hyperbolic equations we shall study later in Section
4.2.6. Although the methods we use here to treat both parabolic and hyperbolic equations
are very similar, we shall see that their solutions behave quite different.
Let us denote by C 2,1 (ΩT ) functions satisfying ut , uxi xj ∈ C(ΩT ). We shall consider
general inequalities of the form
(4.39) ut + Lu ≤ 0 in ΩT
where λ > 0 is a constant, and the coefficients aij , bi , c are all bounded functions in ΩT .
Theorem 4.8. (Weak Maximum Principle) Suppose u ∈ C 2,1 (ΩT ) ∩ C(ΩT ) satisfies
(4.39) and c ≥ 0. Then
Theorem 4.10. (Strong Maximum Principle) Suppose u ∈ C 2,1 (ΩT ) ∩ C(ΩT ) satisfies
(4.39). Let
M = max u = u(x0 , t0 ).
ΩT
Note that for each fixed time t ∈ [0, T ] the operator Lu is uniformly elliptic on Ω.
4.2.1. Weak Solutions. We consider the case that Lu has the divergence form (4.44).
Let B[u, v; t] be the time-dependent bilinear form defined as above.
Definition 4.7. A weak solution to Problem (4.43) is a function u ∈ L2 (0, T ; H01 (Ω))
having weak time-derivatives u0 ∈ L2 (0, T ; L2 (Ω)) and u00 ∈ L2 (0, T ; H −1 (Ω)) such that
(i) hu00 (t), vi + B[u(t), v; t] = (f (t), v) for each v ∈ H01 (Ω) and a.e. time t ∈ [0, T ], and
(ii) u(0) = g, u0 (0) = h. (Note that u ∈ C([0, T ]; L2 (Ω)) and u0 ∈ C([0, T ]; H −1 (Ω)),
and thus u(0) and u0 (0) are well-defined.)
(For instance, we could take {wi } to be the complete set of appropriately normalized eigen-
functions for −∆ in H01 (Ω).)
112 4. Linear Evolution Equations
Fix now a positive integer k. Let Vk be the linear span of {w1 , · · · , wk } and we look for
a function uk : [0, T ] → Vk of the form
k
X
(4.48) uk (t) = di (t)wi ,
i=1
Note that the coefficients αij belong to L∞ (0, T ) and fi ∈ L2 (0, T ). The existence of a
unique solution d ∈ H 2 (0, T ) ⊂ C 1 ([0, T ]) is guaranteed by the (not so) standard existence
theory for ODE (think of approximating αij and fi by smooth functions first and then pass
to limits).
according to the initial data in (4.49) and kuk (0)kH01 (Ω) ≤ kgkH01 (Ω) , we thus obtain
max (ku0k (t)k2L2 (Ω) + A[uk (t), uk (t); t]) ≤ C(khk2L2 (Ω) + kgk2H 1 (Ω) + kf k2L2 (ΩT ) ).
t∈[0,T ] 0
This proves
(4.53) max (kuk (t)kH01 (Ω) + ku0k (t)kL2 (Ω) ) ≤ C(kf kL2 (ΩT ) + kgkH01 (Ω) + khkL2 (Ω) ).
t∈[0,T ]
3. Finally we need to estimate ku00k kL2 (0,T ;H −1 (Ω)) . So, fix any v ∈ H01 (Ω), with kvkH01 (Ω) ≤
1. We write v = v 1 + v 2 , where v 1 ∈ Vk , and (v 2 , wi ) = 0 for all i = 1, 2, · · · , k. (That is, v 2
is in the L2 orthogonal complement of Vk .) Since {wi } are orthogonal in H01 (Ω), we have
kv 1 kH01 (Ω) ≤ kvkH01 (Ω) ≤ 1.
Using (4.49), we have
(u00k (t), v 1 ) + B[uk (t), v 1 ; t] = (f (t), v 1 ).
Then
hu00k (t), vi = (u00k (t), v) = (u00k (t), v 1 ) = (f (t), v 1 ) − B[uk (t), v 1 ; t]
and consequently
|hu00k (t), vi| ≤ C(kf (t)kL2 (Ω) + kuk kH01 (Ω) ).
This implies
ku00k (t)k2H −1 (Ω) ≤ C(kf (t)k2L2 (Ω) + kuk (t)k2H 1 (Ω) ) ∀ t ∈ [0, T ].
0
Proof. 1. According to the energy estimate (4.50), we see that {uk } is bounded in
L2 (0, T ; H01 (Ω)), {u0k } is bounded in L2 (0, T ; L2 (Ω)), and {u00k } is bounded in L2 (0, T ; H −1 (Ω)).
Consequently there exists a subsequence {ukm } of {uk } with km → ∞ and functions
u ∈ L2 (0, T ; H01 (Ω)), with u0 ∈ L2 (0, T ; L2 (Ω)), u00 ∈ L2 (0, T ; H −1 (Ω)), such that
ukm * u
in L2 (0, T ; H01 (Ω)),
(4.54) u0km * u0 in L2 (0, T ; L2 (Ω)),
00
ukm * u00 in L2 (0, T ; H −1 (Ω)).
In fact, by estimate (4.50), we also have u ∈ L∞ (0, T ; H01 (Ω)) ∩ C([0, T ]; L2 (Ω)) and u0 ∈
L∞ (0, T ; L2 (Ω)); moreover, u0 ∈ C([0, T ]; H −1 (Ω)).
2. Fix any integer N and let ψ ∈ C 1 ([0, T ]; H01 (Ω)) have the form
N
X
ψ(t) = ζi (t)wi ,
i=1
where ζi ∈ C 1 ([0, T ]; R). Let k ≥ N , multiply (4.10) by ζi (t), sum i = 1, 2, · · · , N and then
integrate over t ∈ [0, T ] to find
Z T Z T
00
(4.55) huk (t), ψ(t)i + B[uk (t), ψ(t); t] dt = (f (t), ψ(t))dt.
0 0
Now let k = km → ∞ and we have
Z T Z T
00
(4.56) hu (t), ψ(t)i + B[u(t), ψ(t); t] dt = (f (t), ψ(t))dt.
0 0
This equality then holds for all functions ψ ∈ L2 (0, T ; H01 (Ω)), as functions ζ of the given
form are dense in this space. We then take ψ(t) = ζ(t)v with ζ ∈ L2 (0, T ) and v ∈ H01 (Ω)
in (4.56) to obtain
Z T Z T
ζ(t) hu00 (t), vi + B[u(t), v; t] dt =
ζ(t)(f (t), v)dt.
0 0
This holding for all ζ ∈ L2 (0, T ) yields that
(4.57) hu00 (t), vi + B[u(t), v; t] = (f (t), v) ∀ v ∈ H01 (Ω), a.e. t ∈ [0, T ].
3. We need to show the initial data u(0) = g, u0 (0) = h. In (4.55), (4.56), we take
ψ(t) = α(t)v + β(t)w with α, β ∈ C 2 [0, T ] and v, w ∈ H01 (Ω) arbitrarily given such that
ψ(T ) = ψ 0 (T ) = 0, ψ(0) = v and ψ 0 (0) = w. Note that
Z T Z T
hu00k (t), ψ(t)idt = −(u0k (0), v) + (uk (0), w) + (uk (t), ψ 00 (t)) dt,
0 0
(uk (0), w) → (g, w), (u0k (0), v) → (h, v) as k = km → ∞, and
Z T Z T
00 0
hu (t), ψ(t)idt = −hu (0), vi + (u(0), w) + (u(t), ψ 00 (t)) dt,
0 0
Hence, in (4.55), let k = km → ∞, we eventually obtain
−hu0 (0), vi + (u(0), w) = −(h, v) + (g, w) ∀ v, w ∈ H01 (Ω);
hence u(0) = g and u0 (0) = h.
4.2. Second-order Hyperbolic Equations 115
Proof. 1. It suffices to prove that a weak solution u with f = g = 0 must be zero. Unlike
the parabolic case, the proof here is tricky because we cannot insert v = u0 (t) in (4.49) since
u0 (t) ∈
/ H01 (Ω). We instead consider, for each fixed s ∈ [0, T ], the function
(R s
t u(τ )dτ if t ∈ [0, s],
v(t) =
0 if t ∈ [s, T ].
Then for each t ∈ [0, T ], v(t) ∈ H01 (Ω), and so, by the weak solution definition,
Z s
(hu00 (t), v(t)i + B[u(t), v(t); t]) dt = 0.
0
Since u0 (0) = v(s) = 0, we obtain by integration by parts
Z s
(4.58) (−hu0 (t), v 0 (t)i + B[u(t), v(t); t]) dt = 0.
0
As above, we write B[u, v; t] = A[u, v; t] + C[u, v; t]. Note, for all u, v ∈ H01 (Ω) and t ∈ [0, T ],
Z Xn Z X n
C[u, v; t] = bi Di uv + cuv dx = − (Di bi uv + bi uDi v) + cuv dx.
Ω i=1 Ω i=1
Hence
Z s
1 d
ku(t)k2L2 (Ω) − A[v(t), v(t); t] dt
2 0 dt
(4.59) Z s n
1 s
Z Z X
=− C[u(t), v(t); t]dt − a0ij Di vDj vdxdt,
0 2 0 Ω
i,j=1
and consequently,
Z s Z sZ n
X
ku(s)k2L2 (Ω) + A[v(0), v(0); 0] = −2 C[u(t), v(t); t]dt − a0ij Di vDj vdxdt
0 0 Ω i,j=1
Z s
≤C (kv(t)k2H 1 (Ω) + ku(t)k2L2 (Ω) ) dt.
0
0
By Gårding’s inequality for A[u, u; t], we obtain
Z s
2 2
(4.60) ku(s)kL2 (Ω) + kv(0kH 1 (Ω) ≤ C (kv(t)k2H 1 (Ω) + ku(t)k2L2 (Ω) ) dt.
0 0
0
2. Now let Z t
w(t) = u(τ )dτ (t ∈ [0, T ]).
0
116 4. Linear Evolution Equations
Then v(0) = w(s) and v(t) = w(s) − w(t), and hence (4.60) becomes
Z s
2 2
ku(s)kL2 (Ω) + kw(s)kH 1 (Ω) ≤ C (kw(t) − w(s)k2H 1 (Ω) + ku(t)k2L2 (Ω) )dt.
0 0
0
But
kw(t) − w(s)k2H 1 (Ω) ≤ 2kw(t)k2H 1 (Ω) + 2kw(s)k2H 1 (Ω) ,
0 0 0
so we have
Z s
ku(s)k2L2 (Ω) + (1 − 2sC1 )kw(s)k2H 1 (Ω) ≤ C1 (kw(t)k2H 1 (Ω) + ku(t)k2L2 (Ω) )dt.
0 0
0
4.2.5. Regularity. We now study the regularity of weak solutions when the initial data
and coefficients are more regular. Our eventual goal is to prove that the weak solution
is smooth, as long as the coefficients and initial data and the domain are all smooth.
Although the methods and results are similar to the ones used for the regularity study of
parabolic equations as above, as we shall see later, there are some quite essential differences
of regularity concerning these two classes of evolution equations.
We assume the coefficients aij , bi , c are smooth on ΩT . As usual, we assume the uniform
hyperbolicity.
Theorem 4.15. (Improved regularity) (i) Assume g ∈ H01 (Ω), h ∈ L2 (Ω), f ∈ L2 (ΩT ).
Suppose u is the weak solution of (4.43). Then
u ∈ L∞ (0, T ; H01 (Ω)), u0 ∈ L∞ (0, T ; L2 (Ω)),
with the estimate
(4.61) kukL∞ (0,T ;H01 (Ω)) + ku0 kL∞ (0,T ;L2 (Ω)) ≤ C(kgkH01 (Ω) + khkL2 (Ω) + kf kL2 (ΩT ) ),
where C depends only on Ω, T and the coefficients of L.
(ii) If, in addition, g ∈ H 2 (Ω), h ∈ H01 (Ω), f 0 ∈ L2 (0, T ; L2 (Ω)), then
u ∈ L∞ (0, T ; H 2 (Ω)), u0 ∈ L∞ (0, T ; H01 (Ω)),
u00 ∈ L∞ (0, T ; L2 (Ω)), u000 ∈ L2 (0, T ; H −1 (Ω)),
with the estimate
kukL∞ (0,T ;H 2 (Ω)) + ku0 kL∞ (0,T ;H01 (Ω)) + ku00 kL∞ (0,T ;L2 (Ω))
(4.62)
+ ku000 kL2 (0,T ;H −1 (Ω)) ≤ C(kgkH 2 (Ω) + khkH01 (Ω) + kf kH 1 (0,T ;L2 (Ω) ),
where C depends only on Ω, T and the coefficients of L.
Proof. Let {uk } be the Galerkin approximations satisfying (4.49) constructed as above
with {wi } being the complete collection of eigenfunctions with eigenvalues {λi } for −∆ on
H01 (Ω). As before, we assume {wi } is orthogonal on H01 (Ω) and orthonormal on L2 (Ω).
4.2. Second-order Hyperbolic Equations 117
By the uniqueness theorem, the weak solution is obtained as the limit of the Galerkin
approximations {uk }. We prove the theorem by deriving the same estimates for these
approximate solutions independent of k.
1. By energy estimates (4.50), we have
(4.63) max (kuk (t)kH01 (Ω) + ku0k (t)kL2 (Ω) ) ≤ C(kf kL2 (ΩT ) + kgkH01 (Ω) + khkL2 (Ω) )
t∈[0,T ]
+ (f 0 (t), ũ0k (t)) − B̃[uk (t), ũ0k (t); t] − C[ũk (t), ũ0k (t); t].
Notice that, for u ∈ H01 (Ω) ∩ H 2 (Ω), v ∈ H01 (Ω), integration by parts yields
Z X n n
X
B̃[u, v; t] = v (−Dj a0ij Di u − a0ij Dij u) + b0i (x, t)Di u + c0 (x, t)u dx,
Ω i,j=1 i=1
and, hence, for a.e. t ∈ [0, T ] and all u ∈ H 2 (Ω), v ∈ H01 (Ω),
|B̃[u, v; t]| ≤ C(kuk2H 2 (Ω) + kvk2L2 (Ω) ).
118 4. Linear Evolution Equations
≤ C(kũ0k (t)k2L2 (Ω) + A[ũk (t), ũk (t); t] + kf (t)k2L2 (Ω) + kuk (t)k2L2 (Ω) + kf 0 (t)k2L2 (Ω) ).
Hence Gronwall’s inequality implies
kũ0k (t)k2L2 (Ω) + A[ũk (t), ũk (t); t]
Z T
0 2 0
≤ C kũk (0)kL2 (Ω) + A[ũk (0), ũk (0); 0]) + (kf (t)k2L2 (Ω) + kuk (t)k2L2 (Ω) + kf (t)k2L2 (Ω) )dt
0
≤ C(kũ0k (0)k2L2 (Ω) + kũk (0)k2H 1 (Ω) + kf k2H 1 (0,T ;L2 (Ω)) + kgk2H 1 (Ω) + khk2L2 (Ω) ).
0 0
Finally as above, kuk (0)kH 2 (Ω) ≤ CkgkH 2 (Ω) , from which and (4.64), we deduce
sup (ku00k (t)k2L2 (Ω) + kuk (t)k2H 2 (Ω) + ku0k (t)k2H 1 (Ω) )
0
t∈[0,T ]
(4.68)
≤ C(kf k2H 1 (0,T ;L2 (Ω)) + kgk2H 1 (Ω) + khk2H 1 (Ω) ).
0
4. As in the earlier proof for the parabolic equations, we can deduce the estimate for
u000 ∈ L2 (0, T ; H −1 (Ω)) in terms of the right-hand side of (4.68).
Theorem 4.16. (Higher regularity) Let m ∈ {0, 1, 2, · · · }. Assume
dk f
g ∈ H m+1 (Ω), h ∈ H m (Ω), ∈ L2 (0, T ; H m−k (Ω)) (k = 0, 1, · · · , m).
dtk
Suppose the following m-th order compatibility conditions hold:
∈ H01 (Ω), h1 := h ∈ H01 (Ω), · · · ,
g0 := g 2l−2
(4.69) g2l := dt2l−2f (0) − Lg2l−2 ∈ H01 (Ω) ( if m = 2l),
d
2l−1
h2l+1 := ddt2l−1f (0) − Lh2l−1 ∈ H01 (Ω) ( if m = 2l + 1).
Then
dk u
∈ L∞ (0, T ; H m+1−k (Ω)) (k = 0, 1, 2, · · · .m + 1),
dtk
with the estimate
(4.70)
m+1 m
!
X dk u X dk f
k k kL∞ (0,T ;H m−k+1 (Ω)) ≤ C k k kL2 (0,T ;H m−k (Ω)) + kgkH m+1 (Ω) + khkH m (Ω) .
dt dt
k=0 k=0
Proof. Again use induction on m and differentiate the equation with respect to t. Details
are referred to Evans’s book.
4.2. Second-order Hyperbolic Equations 119
4.2.6. Propagation of Disturbances. So far our study of hyperbolic equations has much
paralleled our treatment of parabolic equations, using the Galerkin method. However, we
learned that the classical solution to a second-order parabolic equation has the maximum
principle, which implies an infinite propagation speed of initial disturbances for such
equations. We now study a property for second-order hyperbolic equations that is totally
the opposite phenomenon, namely the finite propagation speed of initial disturbances.
For simplicity, we study the operator of nondivergence form
n
X
Lu = − aij (x)Dij u,
i,j=1
where the coefficients aij are smooth, independent of time,aij = aji , and satisfy the usual
uniform ellipticity condition.
Assume q(x) is a continuous function on Rn and smooth in Rn \ {x0 }, satisfying
n
q(x) > 0 in R \ {x0 }, q(x0 ) = 0,
n
(4.71) X
aij (x)Di qDj q ≤ 1 in Rn \ {x0 }.
i,j=1
We compute e0 (t). In order to do so, note that if f (x, t) is continuous in x and smooth in t
then Z Z Z
d f (x, t)
f (x, t)dx = f (x, t)dx − dS,
dt Kt Kt ∂Kt |∇q(x)|
according to the co-area formula.
2. Therefore, we compute
Z n Z n
X 1 X 1
e0 (t) = ut utt + aij Di uDj ut dx − u2t + aij Di uDj u dS
Kt 2 ∂Kt |∇q|
i,j=1 i,j=1
:= A − B.
Using aij Di uDj ut = Dj (aij ut Di u) − ut Dj (aij Di u) and integration by parts, we have
Z X n Z Xn
A= ut utt − Dj (aij Di u) dx + aij ν j ut Di udS
Kt i,j=1 ∂Kt i,j=1
(4.72) Z n Z n
X X
=− ut Dj aij Di u dx + aij ν j ut Di udS,
Kt i,j=1 ∂Kt i,j=1
an inner product on Rn , with norm kξk = hξ, ξi1/2 ; hence, the Cauchy-Schwarz inequality
|hξ, ηi| ≤ kξkkηk implies
1/2 1/2
Xn Xn n
X
aij (x)ν j Di u ≤ aij (x)Di uDj u aij (x)ν i ν j
i,j=1 i,j=1 i,j=1
1/2 1/2
n n
X X Di qDj q
= aij (x)Di uDj u aij (x)
|∇q|2
i,j=1 i,j=1
1/2
n
X 1
≤ aij (x)Di uDj u (x ∈ ∂Kt ),
|∇q|
i,j=1
= Ce(t) + B.
3. Therefore, we deduce
e0 (t) = A − B ≤ Ce(t) + B − B = Ce(t) (0 < t < t0 ).
Since e(0) = 0 and e(t) ≥ 0, we deduce from Gronwall’s inequality, that e(t) ≡ 0 for all
0 ≤ t ≤ t0 . This proves u ≡ 0 in K.
4.3. Hyperbolic Systems of First-order Equations 121
4.3.1. Notations and Definitions. Consider systems of linear first-order PDE having
the form
n
X
(4.73) ut + Bj (x, t)Dj u = f in Rn × (0, ∞),
j=1
subject to the initial condition
(4.74) u=g on Rn × {t = 0}.
The unknown is u : Rn × [0, ∞) → Rm , u = (u1 , · · · , um ), and the functions Bj : Rn ×
[0, ∞) → Mm×m (j = 1, 2, · · · , n), f : Rn × [0, ∞) → Rm and g : Rn → Rm are given,
Definition 4.9. (i) The system of PDE (4.73) is called a hyperbolic system if the m × m
matrix X
B(x, t; ξ) = ξj Bj (x, t)
j=1
is diagonalizable for each x, ξ ∈ Rn , t ≥ 0. In other words, (4.73) is a hyperbolic system
if for each x, ξ, t the matrix B(x, t; y) defined above has m real eigenvalues
λ1 (x, t; ξ) ≤ λ2 (x, t; ξ) ≤ · · · ≤ λm (x, t; ξ)
and corresponding eigenvectors {rk (x, t; ξ)}m m
k=1 that form a basis of R .
(ii) We say (4.73) is a symmetric hyperbolic system if Bj (x, t) is symmetric for each
x ∈ Rn , t ≥ 0.
(iii) The system is called strictly hyperbolic if if for each x, ξ, t the matrix B(x, t; ξ)
defined above has m distinct real eigenvalues
λ1 (x, t; ξ) < λ2 (x, t; ξ) < · · · < λm (x, t; ξ).
4.3.2. Vanishing Viscosity Method. We study the initial value problem (4.73),(4.74),
with
Bj ∈ C 2 (Rn × [0, T ]; Mm×m ) is symmetric,
2
(4.75) sup (|Bj | + |Dx,t Bj | + |Dx,t Bj |) < ∞,
Rn ×[0,T ]
We shall use the vanishing viscosity method to prove the existence of weak solution. To
this end, we approximate the initial value problem by the parabolic problem
(
ut − ε∆u + nj=1 Bj Dj u = f in Rn × (0, T ],
P
(4.77)
u = g on Rn × {t = 0},
for 0 < ε ≤ 1, gε := ηε ? g. The second-order term −ε∆u is called the viscosity term,
which tends to regularize the original first-order system.
Theorem 4.19. (Existence of approximate solutions) For each 0 < ε ≤ 1, there exists
a unique solution u = uε of (4.77), with
uε ∈ L2 (0, T ; H 3 (Rn ; Rm )), u0ε ∈ L2 (0, T ; H 1 (Rn ; Rm )).
Proof. 1. Set X = L∞ (0, T ; H 1 (Rn ; Rm )). For each v ∈ X, consider the linear system
(
ut − ε∆u = f − nj=1 Bj Dj v in Rn × (0, T ],
P
(4.78)
u = g on Rn × {t = 0}.
The right-hand side is bounded in L2 , there exists a unique solution u ∈ L2 (0, T ; H 2 (Rn ; Rm )),
with u0 ∈ L2 (0, T ; L2 (Rn ; Rm )). This solution u can be expressed by the Duhamel formula
using the heat kernel. From this we can also show that u ∈ X = L∞ (0, T ; H 1 (Rn ; Rm )).
Hence we define a map S : X → X by setting u = S(v). Let v1 ∈ X and u1 = S(v1 ). Set
w = u − u1 and z = v − v1 . Then
(
wt − ε∆w = − nj=1 Bj Dj z in Rn × (0, T ],
P
w = 0 on Rn × {t = 0}.
From the representation of w in terms of the heat kernel and nj=1 Bj Dj z, we have
P
n
X
kwkL∞ (0,T ;H 1 (Rn ;Rm )) ≤ C(ε)k Bj Dj zkL2 (0,T ;L2 (Rn ;Rm ))
j=1
≤ C(ε)kzkL2 (0,T ;H 1 (Rn ;Rm ))
≤ C(ε)T 1/2 kzkL∞ (0,T ;H 1 (Rn ;Rm )) .
3. If C(ε)T 1/2 < 1 then S is a strict contraction on X; hence it has a unique fixed point
u = uε : S(uε ) = uε . Then u = uε solves (4.77) for such a T > 0. If C(ε)T 1/2 ≥ 1, then
1/2
we choose 0 < T1 < T so that C(ε)T1 < 1 and repeat the above argument on the time
intervals [0, T1 ], [T1 , 2T1 ], etc, to obtain a weak solution u = uε for all T > 0. Finally the
high regularity of such a solution u follows from parabolic regularity theory.
Theorem 4.20. (Energy estimates) There exists a constant C, depending only on n and
the coefficients,such that
max (kuε (t)kH 1 (Rn ;Rm ) + ku0ε (t)kL2 (Rn ;Rm ) )
t∈[0,T ]
(4.80)
≤ C(kgkH 1 (Rn ;Rm ) + kf kL2 (0,T ;H 1 (Rn ;Rm ) + kf 0 kL2 (0,T ;L2 (Rn ;Rm ) )
for all 0 < ε ≤ 1.
4.3. Hyperbolic Systems of First-order Equations 123
Proof. 1. We compute
n
d 1 X
kuε (t)k2L2 (Rn ;Rm ) = (uε (t), u0ε (t)) = uε , f (t) + ε∆uε (t) − Bj Dj uε (t) .
dt 2
j=1
Note that
(uε (t), ε∆uε (t)) = −εk∇uε (t)k2L2 (Rn ) ≤ 0 (0 < t ≤ T ).
2. Suppose v ∈ C0∞ (Rn ; Rm ). Then, by the symmetry of Bj (this is the only place the
symmetry assumption is used: if B is symmetric then Ba · b = Bb · a),
X n Z X n
B[v, v; t] = v, Bj Dj v = (Bj Dj v) · v dx
j=1 Rn j=1
n Z n Z
1 X 1X
= Dj [(Bj v) · v]dx − [(Dj Bj )v) · v]dx
2 n 2 Rn
j=1 R j=1
n Z
1X
=− [(Dj Bj )v) · v]dx.
2 Rn
j=1
We therefore deduce
d 1
kuε (t)k2L2 (Rn ;Rm ) ≤ C kuε (t)k2L2 (Rn ;Rm ) + kf (t)k2L2 (Rn ;Rm ) .
dt 2
So Gronwall’s inequality and kgε kL2 ≤ kgkL2 will yield
(4.82) max kuε (t)k2L2 (Rn ;Rm ) ≤ C kgk2L2 (Rn ;Rm ) + kf k2L2 (0,T ;L2 (Rn ;Rm )) .
t∈[0,T ]
Proof. This follows from the energy estimates above in exactly the same fashion as before.
Theorem 4.22. (Uniqueness of weak solution) A weak solution to problem (4.77) is
unique.
Proof. It suffices to show that the only weak solution to (4.77) with f ≡ g ≡ 0 is u ≡ 0.
To verify this, note that (u0 (t), u(t)) + B[u(t), u(t); t] = 0 for a.e. t ∈ [0, T ] and, by (4.81),
|B[u(t), u(t); t]| ≤ Cku(t)k2L2 (Rn ;Rm ) ,
so we have
d
ku(t)k2L2 (Rn ;Rm ) ≤ Cku(t)k2L2 (Rn ;Rm ) ,
dt
whence Gronwall’s inequality forces u ≡ 0.
since s > n2 + m. Therefore ut exists and is continuous on Rn × [0, ∞). A similar argument
shows that Dj u exists and is continuous for all j = 1, 2, · · · , n (the proof is similar since ξ
is like B(ξ)). Furthermore, we can differentiate inside the integral in (4.89) to confirm that
u solves the system (4.86).
Effectively, T (t) maps the initial data u0 into the current value of the solution u(t).
Before we discuss the general theory for (4.90) with A being a linear operator defined
in a subspace of a Banach space, we review some definitions and elementary properties.
Example 4.24. Let L : D(L) ⊂ L2 (0, 1) → L2 (0, 1) be the differential operator L = d/dx,
where D(L) = H01 (0, 1). To show that L is closed, let un → u, u0n → f in L2 (0, 1), where
un ∈ D(L). Passing to the limit in
Z 1 Z 1
un v 0 dx = − u0n vdx for all v ∈ C0∞ (0, 1)
0 0
we see, by the definition of weak derivative, that u0 = f and un → u in H 1 (0, 1). Since
H01 (0, 1) is closed, we have u ∈ H01 (0, 1).
Lu = ∆u, u ∈ D(L).
Note that we are considering L as an operator on L2 (Ω). Clearly L is densely defined. It is
unbounded, for if we consider {ϕn }, the sequence of eigenfunctions of −∆, then kϕn k2 = 1
while kLϕn k2 = λn → ∞ as n → ∞.
To see that L : D(L) ⊂ L2 (Ω) → L2 (Ω) is a closed operator, let un ∈ D(L) with
un → u, Lun → f . Applying the estimate kuk2,2 ≤ ckLuk2 to un − um , it follows that {un }
is a Cauchy sequence in H 2 (Ω) and thus kun − vk2,2 → 0 for some v ∈ H 2 (Ω). Clearly u = v
and u ∈ D(L). Since L : H 2 (Ω) → L2 (Ω) is continuous, Lun → Lu which yields Lu = f .
Hence L is closed.
Theorem 4.26. (Closed Graph Theorem) If X, Y are Banach spaces and if T : D(T ) ⊂
X → Y is a closed linear operator with closed domain, then T is bounded.
Definition 4.11. Let X be a Banach space and let T : D(T ) ⊂ X → X be a closed operator
on X.
(i) We say a real number λ belongs to the resolvent set ρ(T ) of T , provided the
operator
λI − T : D(T ) → X
is one-to-one and onto.
(ii) If λ ∈ ρ(T ), the resolvent operator Rλ : X → X is defined by
Rλ u := (λI − T )−1 u (u ∈ X);
that is, Rλ = (λI − T )−1 .
Remark 4.12. (i) Note that, since T is closed, by (4.27), Rλ is a bounded linear operator
on X if λ ∈ ρ(T ). Furthermore,
T Rλ u = Rλ T u (u ∈ D(T )).
Definition 4.13. Let X be a Banach space. A family {T (t)} ⊂ B(X) (0 ≤ t < ∞) is called
a strongly continuous semigroup of operators if
(i) T (t)T (s) = T (t + s), t, s ≥ 0 (semigroup property)
(ii) T (0) = I
(iii) For all u ∈ X, T (t)u is strongly continuous in t ∈ [0, ∞), i.e.,
kT (t + h)u − T (t)uk → 0 as h → 0.
For simplicity we say that T (t) is a C0 semigroup.
Moreover, for fixed u ∈ X, we write T (·)u ∈ C([0, ∞); X). If in addition the map
t → T (t) is continuous in the uniform operator topology, i.e., kT (t + h) − T (t)k → 0 for
t, t + h ≥ 0, then the family {T (t)} is called a uniformly continuous semigroup. If
the C0 semigroup {T (t)} satisfies the property kT (t)k ≤ 1 for t ≥ 0, then it is called a
contraction semigroup.
Note that T (t) and T (s) commute as a consequence of (i), and that T (t)u is also strongly
continuous in u for each fixed t ≥ 0. (kT (t)u − T (t)vk ≤ kT (t)kku − vk.)
Example 4.28. Let A ∈ B(X) where X is a Banach space. Then the series ∞ n n
P
n=0 (A /n!)t
converges in the uniform operator topology for any real number t. In fact, set
n
X
Sn = (Ak /k!)tk
k=0
Thus {Sn } converges to a bounded linear operator, in the uniform operator topology, which
we denote by etA . From the above estimate we see that
ketA k ≤ e|t|kAk .
Let u(t) = etA u0 where u0 ∈ X. Since e(t+s)A = etA esA , it follows that
hA
u(t + h) − u(t) e −I
− Au = − A u.
h h
However,
∞
ehA − I X
k − Ak ≤ (kAkk /k!)|h|k−1
h
k=2
e|h|kAk −1
= − kAk → 0 as h → 0.
|h|
Hence u0 (t) exists for all t and equals Au, and so we have shown that u(t) = etA u0 satisfies
the Cauchy problem
du
= Au (t > 0), u(0) = u0 .
dt
4.4. Semigroup Theory 129
Proof. In fact, since the function g(t) = kT (t)uk is continuous on [0, 1] for each fixed
u ∈ X, we have supt∈[0,1] kT (t)uk < ∞. Hence, by the Uniform Boundedness Principle,
there is a constant M > 0 such that kT (t)k ≤ M for t ∈ [0, 1]. Let ω = log M . Then ω ≥ 0,
since M ≥ 1 by virtue of T (0) = I. Now let t be given and let n be the least integer greater
than or equal to t. By virtue of the semigroup property (T (t/n))n = T (t) and thus
kT (t)k = k(T (t/n))n k ≤ M n ≤ M t+1 = M eωt .
Proof. (a) This follows directly from the definition since Ah is linear.
(b) Since T (t) and T (h) commute and kT (t)k < ∞, we have
Ah T (t)u = T (t)Ah u → T (t)Au, as h → 0+ .
Hence
T (t)u ∈ D(A), AT (t)u = T (t)Au = D+ T (t)u.
Next we consider D− T (t)u if t > 0. Note that
T (t)u − T (t − h)u T (h)u − u
− T (t)Au = T (t − h) − Au
h h
+ (T (t − h) − T (t))Au.
But
T (h)u − u T (h)u − u
kT (t − h) − Au k ≤ M eωt k − Auk → 0
h h
as h → 0+ and
k(T (t − h) − T (t))Auk → 0 as h → 0+
which implies the desired result.
(c) The abstract function T (t)u is differentiable by (b) and its derivative T (t)Au is
continuous in t. The conclusion follows by integrating (4.92).
(d)
Z t+h Z t+h
−1 −1
kh f (s)T (s)u ds − f (t)T (t)uk = kh (f (s)T (s)u − f (t)T (t)u) dsk
t t
Z t+h Z t+h
−1 −1
≤h |f (s)|k(T (s)u − T (t)uk ds + h kT (t)uk|f (s) − f (t)| ds.
t t
Since the functions T (t)u and f (t) are continuous in t, by choosing h small enough so that
k(T (s)u − T (t)uk < ε and |f (s) − f (t)| < ε for |t − s| < h, the result easily follows.
(e) Let h > 0 and consider
Z t t
T (h) − I
Z
1
T (s)u ds = (T (s + h)u − T (s)u)ds
h 0 h 0
Z t+h Z h
1
= T (s)u ds − T (s)u ds
h t 0
→ T (t)u − u
Rt
by (d). Thus T (s)u ds ∈ D(A) and (e) follows by the definition of A.
0
Rh Rh
(f) Let u ∈ X. Then 0 T (s)u ds ∈ D(A) and by (d), limh→0 h−1 0 T (s)u ds =
T (0)u = u. Thus D(A) = X. Let un ∈ D(A) with un → u, Aun → v in X. Now
Z h
T (h)un − un −1
Ah u = lim = lim h T (s)Aun ds
n→∞ h n→∞ 0
where the last term follows from (c). But Aun → v and so
Z h
−1
Ah u = h T (s)v ds → v as h → 0
0
by virtue of (d). Thus u ∈ D(A) and Au = v.
4.4. Semigroup Theory 131
Remark 4.15. Since the map t → T (t)Au is continuous, it follows from (b) that T (t)u is
continuously differentiable from [0, ∞) with values in X. Also T (t)u ∈ D(A) as proved, so
it has values in D(A) as well. Further, the continuous differentiability into X also proves
the continuity into D(A) (with the graph norm).
Proof. The existence is a consequence of Theorem 4.31 (b). To prove uniqueness, let v(t)
be any solution of the Cauchy problem and set F (s) = T (t − s)v(s). Then
F (s + h) − F (s) = T (t − s − h) (v(s + h) − v(s)) − (T (t − s) − T (t − s − h)) v(s).
Since v(s) ∈ D(A), Theorem 4.31 implies that F (s) is strongly differentiable in s and
(d/ds)F (s) = −AT (t − s)v(s) + T (t − s)v 0 (s)
= −AT (t − s)v(s) + T (t − s)Av(s) = 0, 0 ≤ s ≤ t.
Hence by the mean value theorem, F (s) =constant for 0 ≤ s ≤ t. In particular, F (t) = F (0)
or v(t) = T (t)u0 = u(t).
Remark 4.16. (i) If T (t)u0 is differentiable for every u0 ∈ X and t ≥ 0, then in particular,
(d/dt)(T (t)u0 )|t=0 = Au0 . Hence, D(A) = X and A (being a closed operator) must be
bounded. Thus if A is unbounded, then T (t)u0 is not differentiable for all u0 ∈ X. We can
however consider u(t) = T (t)u0 as a generalized solution of the Cauchy problem.
(ii) From the uniqueness in Theorem 4.32 we have that a linear operator A : D(A) → X
which is densely defined can be the infinitesimal generator of at most one C0 semigroup
{T (t)}. Moreover, if {T (t)} is a C0 semigroup whose infinitesimal generator A is bounded,
then T (t) = eAt since then A is the infinitesimal generator of T (t) and etA .
Hence λRλ u → u as λ → ∞ if u ∈ D(A). But since kλRλ k ≤ 1 and D(A) is dense, we have
lim λRλ u = u (u ∈ X).
λ→∞
So, if u ∈ D(A),
lim Aλ u = lim λARλ u = lim λRλ Au = Au.
λ→∞ λ→∞ λ→∞
4. Define
∞
2 tR
X (λ2 t)k
Tλ (t) := etAλ = e−λt eλ λ
= e−λt Rλk .
k!
k=0
Observe, since kRλ k ≤ 1/λ,
∞
−λt
X (λ2 t)k
kTλ (t)k ≤ e =1 (t ≥ 0).
k!
k=0
Proof.
k(λI − A)ukkuk ≥ Re((λI − A)u, u) = λkuk2 − Re(Au, u) ≥ λkuk2 .
Theorem 4.38. (Lumer-Phillips) Let A : D(A) → H be a densely defined linear operator.
(a) If A is a generator of a C0 contraction semigroup {T (t)}, then A is dissipative
and R(λI − A) = H for all λ > 0.
4.4. Semigroup Theory 135
(b) If A is dissipative and there exists a λ0 > 0 such that R(λ0 I − A) = H, then
A is a generator of a C0 contraction semigroup. In particular, A is maximal
dissipative.
Proof. (a) By the Hille-Yosida Theorem, (0, ∞) ⊂ ρ(A) and therefore λI − A is onto H
for all λ > 0. Furthermore,
|(T (t)u, u)| ≤ kT (t)kkuk2 ≤ kuk2 .
Hence
T (t)u − u 1
Re , u = (Re(T (t)u, u) − kuk2 ) ≤ 0.
t t
+
Let u ∈ D(A) and let t → 0 obtaining Re(Au, u) ≤ 0.
(b) Since R(λ0 I − A) = H, it follows from (4.97) that λ0 ∈ ρ(A) and A is closed. If
R(λI − A) = H for all λ > 0, then (0, ∞) ⊂ ρ(A) and kR(λ; A)k ≤ λ−1 by (4.97). The
desired result then follows from the Hille-Yosida Theorem.
In order to prove that R(λI − A) = H for all λ > 0, consider the open set
Γ = {λ : λ > 0 and R(λI − A) = H}.
Note that λ ∈ Γ implies λ ∈ ρ(A), and since ρ(A) is open, there is a neighborhood of λ
whose intersection with the real line is in Γ. By hypothesis, λ0 ∈ Γ. Hence if Γ is closed in
(0, ∞), then Γ = (0, ∞). Let λn → λ > 0, λn ∈ Γ. For every v ∈ H, there exist un ’s such
that
(4.98) λn un − Aun = v for all n.
From (4.97) it follows that kun k ≤ λ−1
n kvk ≤ c. Now by (4.97) again,
Proof. In view of (b) of Theorem 4.38, it suffices to show that R(I − A) = H. Since A is
closed so is I − A. Moreover, (4.97) and Corollary 4.27 imply R(I − A) is a closed subspace
of H and therefore R(I − A) = (N (I − A*))⊥ . Consequently, it suffices to show that I −A*
is one-to-one. But this follows from the fact that A* is dissipative.
While we could solve the initial value problem in [0, ∞) when u0 ∈ D(A) and not for
general u0 ∈ H, one can show that if A is self-adjoint then we can solve the problem for
any initial data u0 ∈ H. The price we pay for this is the lack of differentiability at t = 0.
Theorem 4.40. Let A be a self-adjoint maximal dissipative operator and let u0 ∈ H. Then
there exists a unique u such that
u ∈ C([0, ∞); H) ∩ C 1 ((0, ∞); H) ∩ C((0, ∞); D(A))
and u satisfies the initial value problem (4.93). (Continuity in D(A) is with respect to the
graph norm.)
136 4. Linear Evolution Equations
Proof. We only prove uniqueness. Let u1 , u2 be two solutions of (4.93). Set v(t) = ku1 (t)−
u2 (t)k2 which is continuous and such that v(0) = 0. Taking the inner product of u0 = Au
with u1 (t) − u2 (t) for u = u1 and u = u2 and using the dissipativity of A, we get
d
v(t) ≤ 0
dt
whence it follows that v(t) ≡ 0 and so u1 (t) = u2 (t) for t ≥ 0.
which a special case studied above. We assume L has the divergence structure, satisfies the
uniform ellipticity condition, and has coefficients that are smooth and independent of time
t. We assume domain Ω is bounded and has smooth boundary ∂Ω.
We propose problem (4.99) as the flow determined by a semigroup on X = L2 (Ω). For
this we set
D(A) = H01 (Ω) ∩ H 2 (Ω)
and define
Au := −Lu (u ∈ D(A)).
Recall the Gårding’s inequality
(4.100) βkuk2H 1 (Ω) ≤ B[u, u] + γku2L2 (Ω) .
0
Moreover, u ∈ H01 (Ω). (Trace operator is continuous from H 1 (Ω) to L2 (∂Ω).) Therefore
u ∈ D(A). Furthermore the strong convergence uk → u in H 2 (Ω) implies Auk → Au in
L2 (Ω), and thus f = Au. By definition, A is closed.
3. We next prove the condition (γ, ∞) ⊂ ρ(A); that is, λI − A is one-to-one and onto
for all λ > γ. By Lax-Milgram’s Theorem, for all λ ≥ γ, the BVP
(
Lu + λu = f in Ω,
(4.101)
u=0 on ∂Ω
has a unique weak solution u ∈ H01 (Ω) for each f ∈ L2 (Ω). The global regularity theory
shows that u ∈ D(A), and
λu − Au = f.
Thus (λI − A) : D(A) → X is one-to-one and onto, for all λ ≥ γ. This proves [γ, ∞) ⊂ ρ(A).
4. Let Rλ = R(λ, A) = (λI − A)−1 . We will show
1
kRλ k ≤ (λ > γ)
λ−γ
as required for generating a C0 γ-contraction semigroup. To show this, consider the weak
form of problem (4.101):
B[u, v] + λ(u, v) = (f, v) ∀ v ∈ H01 (Ω),
where, as usual, (, ) stands for the L2 -inner product. Set v = u and recall the Gårding’s
inequality to compute
(λ − γ)kuk2L2 (Ω) ≤ kf kL2 (Ω) kukL2 (Ω) .
This implies, as u = Rλ f ,
1
kRλ f kL2 (Ω) = kukL2 (Ω) ≤ kf kL2 (Ω) .
λ−γ
This bound is valid for all f ∈ L2 (Ω), which proves the desired claim.
Example 4.42. We study the heat equation:
ut (x, t) − ∆u(x, t) = 0 in Ω × [0, ∞),
(4.102) u(x, t) = 0 on ∂Ω × [0, ∞),
u(x, 0) = u0 (x), x ∈ Ω.
Theorem 4.43. Let u0 ∈ L2 (Ω). Then there exists a unique solution u of (4.102) such that
u ∈ C([0, ∞); L2 (Ω)) ∩ C 1 ((0, ∞); L2 (Ω)) ∩ C((0, ∞); H 2 (Ω) ∩ H01 (Ω)).
we have that A is dissipative. It is maximal dissipative; that is, R(I − A) = H, for, by the
Lax-Milgram, there exists a unique u ∈ H01 (Ω) such that
(I − A)u = u − ∆u = f for all f ∈ L2 (Ω).
Also by elliptic regularity, u ∈ H 2 (Ω) and so u ∈ D(A); hence R(I − A) = H. Finally, we
saw earlier that A is self-adjoint. Hence we can apply Theorem 4.40 to deduce the desired
conclusions.
Remark 4.19. Note that however badly behaved u0 ∈ L2 (Ω) may be, u(x, t) is very smooth
for all t > 0. This is known as the strong regularizing effect of the heat operator. In
particular, this shows that the heat equation is irreversible in time, i.e., we cannot always
solve the problem
ut (x, t) − ∆u(x, t) = 0 in Ω × [0, T ),
u(x, t) = 0 on ∂Ω × (0, T ),
u(x, T ) = uT (x), x ∈ Ω.
which a special case studied above. We assume domain Ω is bounded and has smooth
boundary ∂Ω and L has the symmetric form:
Xn
Lu = − Dj (aij (x)Di u) + c(x)u,
i,j=1
where aij (x) = aji (x) and c(x) ≥ 0 are all smooth functions and (aij (x)) ≥ θI. Hence
(4.104) B[u, u] ≥ θkuk2H 1 (Ω) (u ∈ H01 (Ω)).
0
Now take Hilbert space X = H01 (Ω) × L2 (Ω), with the inner product
hh(u, v), (f, g)ii = B[u, f ] + (v, g)L2 (Ω)
and the norm
k(u, v)k := (B[u, u] + kvk2L2 (Ω) )1/2 .
Define
D(A) := [H 2 (Ω) ∩ H01 (Ω)] × H01 (Ω)
and define
A(u, v) := (v, −Lu) ∀ (u, v) ∈ D(A).
Theorem 4.44. (Second-order hyperbolic PDE as semigroup) The operator A gen-
erates a C0 contraction semigroup {T (t)}t≥0 on H01 (Ω) × L2 (Ω). Therefore (u(t), v(t)) =
T (t)(g, h) defines the solution to (4.105) for any given (g, h) ∈ [H01 (Ω) ∩ H 2 (Ω)] × H01 (Ω).
4.4. Semigroup Theory 139
Proof. We must verify the hypotheses of the Hille-Yosida-Phillips Theorem. Therefore the
evolution problem again becomes a study of the spectral properties of the linear operator
A defined above.
1. D(A) is clearly dense in L2 (Ω).
2. We show A is closed. Indeed, let {(uk , vk )} be a sequence in D(A) with
Putting v = λu − f, we obtain
λ(kvk2L2 (Ω) + B[u, u]) = (v, g)L2 (Ω) + B[u, f ] = hh(u, v), (f, g)ii ≤ k(u, v)kk(f, g)k.
This implies
1
k(u, v)k ≤ k(f, g)k;
λ
so
1
kRλ (f, g)k = k(u, v)k ≤ k(f, g)k.
λ
1
This bound is valid for all (f, g) ∈ X, which proves the desired claim kRλ k ≤ λ for all
λ > 0.
140 4. Linear Evolution Equations
Since T (t) is bounded for each t ≥ 0 and f (s) is continuous, the above integrals exist. Now
Z t+h Z t
[v(t + h) − v(t)]/h = h−1 T (s)f (t + h − s)ds − h−1 T (s)f (t − s)ds
0 0
Z t
= T (s)[f (t + h − s) − f (t − s)]/h ds
0
Z t+h
−1
+ h T (s)f (t + h − s) ds.
t
it follows that Z t Z t
lim Ah T (t − s)f (s) ds = A T (t − s)f (s) ds
h→0 0 0
which yields
Z t
v 0 (t) = A T (t − s)f (s) ds + f (t).
0
4.4. Semigroup Theory 141
As a result we get
Z t
du
= AT (t)u0 + A T (t − s)f (s) ds + f (t)
dt 0
= Au(t) + f (t)
and the proof is complete.
Remark 4.20. The expression (4.108) is called the variation of constants or Duhamel
formula. If the function f is integrable, then (4.108) still makes sense. Then u defined
by that formula is called a generalized solution or mild solution of (4.107). It can be
shown that a generalized solution always exists, but need not be a classical solution.
Index
143
144 Index
sequence
weakly convergent, 7
set
convex, 1
totally bounded, 45
weakly closed, 17
simple, 52
Smoothness of weak solution Theorem, 109, 119
Sobolev conjugate, 40
Sobolev Inequalities Theorem, 39
Sobolev space of order k and power p, 25
spectrum, 10
strictly hyperbolic, 121
Strong Maximum Principle, 69
strong maximum principle, 110
Strong Maximum Principle Theorem, 110
strong solution, 73
strongly measurable, 53
structural conditions, 56
symmetric hyperbolic system, 121
system of second-order (quasilinear) partial
differential equations in divergence form, 56
Trace Theorem, 37
Triangle inequality, 2, 4