Solution to Partial Differential Equations: Methods and Applications (Robert McOwen) Section 6.1
- If X is a normed vector space, prove that |‖x‖−‖y‖|≤‖x−y‖.
- Show that the norm defines a continuous function on X.
- If X is a real inner product space, prove (1).
- By the triangle inequality, we have
‖x‖−‖y‖=‖(x−y)+y‖−‖y‖≤‖x−y‖+‖y‖−‖y‖=‖x−y‖,‖y‖−‖x‖=‖(y−x)+x‖−‖x‖≤‖y−x‖+‖x‖−‖x‖=‖x−y‖.
Thus, we obtain −‖x−y‖≤‖x‖−‖y‖≤‖x−y‖, i.e., |‖x‖−‖y‖|≤‖x−y‖. - Let f:X→R be defined by f(x)=‖x‖ for x∈X. For any ϵ>0, we pick δ=ϵ. Then for x,y∈X and ‖x−y‖<δ, we use (a) to find
|f(x)−f(y)|=|‖x‖−‖y‖|≤‖x−y‖<δ=ϵ.
This shows f is continuous on X. - If y=0, then we have ‖y‖=√⟨0,0⟩=0 and
⟨x,0⟩=¯⟨0,x⟩=¯⟨0+0,x⟩=¯⟨0,x⟩+⟨0,x⟩=¯⟨0,x⟩+¯⟨0,x⟩=⟨x,0⟩+⟨x,0⟩,
which shows ⟨x,0⟩=0. Thus, |⟨x,0⟩|=0=‖x‖‖y‖, i.e., the equality holds true (even over the complex field). Now we suppose y≠0 and consider the function g g:R→R be defined by g(t)=‖x−ty‖2. Clearly, g is nonnegative on R. Then we observe that0≤‖x−ty‖2=⟨x−ty,x−ty⟩=⟨x,x−ty⟩−t⟨y,x−ty⟩=⟨x−ty,x⟩−t⟨x−ty,y⟩(since$X$isarealinnerproductspace)=⟨x,x⟩−t⟨y,x⟩−t⟨x,y⟩+t2⟨y,y⟩=‖x‖2−2t⟨x,y⟩+t2‖y‖2for t∈R.
Taking ⟨x,y⟩‖y‖2, we get‖x‖2−2⟨x,y⟩2‖y‖2+⟨x,y⟩2‖y‖4⋅‖y‖2≥0,
which implies ⟨x,y⟩2≤‖x‖2‖y‖2, i.e., |⟨x,y⟩|≤‖x‖‖y‖. The proof is complete. - Use Hölder's inequality to prove the generalized Hölder's inequality
∫Ω|uvw|dx≤‖u‖p‖v‖q‖w‖r
for u∈Lp(Ω), v∈Lq(Ω), and w∈Lr(Ω), where p−1+q−1+r−1=1. - Use Hölder's inequality to prove
‖u‖q≤‖u‖λp‖u‖1−λrfor u∈Lp(Ω)∩Lr(Ω),
where p≤q≤r and q−1=λp−1+(1−λ)r−1. - let ℓ2 denote the space of all sequences of real numbers {an}∞n=1 such that ∑a2n<∞.
- Verify that ⟨{an},{bn}⟩=∑anbn is an inner product on ℓ2.
- Verify that ℓ2 is a Hilbert space (i.e., complete in the induced norm).
- Let 0={0}∞n=1. Then ⟨0,0⟩=∑02=0. For {an}≠0, i.e., there exists m∈N such that am≠0, we have
⟨{an},{an}⟩=∑a2n≥a2m>0.
Since {an} and {bn} are sequences of real numbers, it is clear that⟨{an},{bn}⟩=∑anbn=¯∑bnan=¯⟨{bn},{an}⟩.
For λ,μ∈R and {an},{bn},{cn}∈ℓ2, we have⟨λ{an}+μ{bn},{cn}⟩=∑(λan+μbn)cn=λ∑ancn+μ∑bncn=λ⟨{an},{cn}⟩+μ⟨{bn},{cn}⟩.
Therefore, ⟨⋅,⋅⟩ is an inner product on ℓ2. - Let {an}(k)={a(k)n} be the Cauchy sequence in ℓ2 for k∈N, i.e., for any ϵ>0, there exists K1∈N such that ‖{an}(k)−{an}(m)‖<ϵ for k,m>K1, where ‖{an}‖=√⟨{an},{an}⟩. In addition, by the triangle inequality, there exists K2 such that ‖{a(k)n}‖<1+‖{a(K1)n}‖:=M for k>K2. For any n∈N and k,m>N, we have
|a(k)n−a(m)n|≤√(a(k)n−a(m)n)2≤√∞∑j=1(a(k)j−a(m)j)2=‖{an}(k)−{an}(m)‖<ϵ.
Thus, for any fixed n∈N, {a(k)n}∞k=1 is Cauchy sequence in R, i.e., limk→∞a(k)n exists and is denoted by a(∞)n. Now we claim that the sequence {a(∞)n} is in ℓ2. For any positive integer N∈N, we notice thatN∑n=1(a(∞)n)2=N∑n=1limk→∞(a(k)n)2=limk→∞N∑n=1(a(k)n)2≤limk→∞∞∑n=1(a(k)n)2≤limk→∞M2=M2.
Since the upper bound for N∑n=1(a(∞)n)2 is independent of N, the series N∑n=1(a(∞)n)2 is convergent, i.e., ∞∑n=1(a(∞)n)2<∞ and hence {a(∞)n}∈ℓ2. Finally, we note thatN∑n=1(a(k)n−a(∞)n)2=N∑n=1(a(k)n−limm→∞a(m)n)2=limk→∞N∑n=1(a(k)n−a(m)n)2≤limk→∞∞∑n=1(a(k)n−a(m)n)2≤limk→∞‖{a(k)n}−{a(m)n}|2≤ϵ2for N∈N and k≥K1.
Therefore, ‖{a(k)n−a(∞)n‖≤ϵ for k≥K1, which implies limk→∞‖{a(k)n−a(∞)n‖=0, i.e., a(k)n converges to a(∞)n in ℓ2. Therefore, ℓ2 is a Hilbert space. - Let Ω=(0,1).
- Verify that du/dx (given in Example 4) is the weak derivative of u.
- For what values of α is u(x)=|x|α in H1,2(Ω)?
- For ϕ∈C∞0(Ω)=C∞0((0,1)), it is easy to verify that ∫Ωudϕdxdx=−∫Ωϕdudxdx as follows:
∫20udϕdxdx=∫1/20xdϕdxdx+∫11/2(1−x)dϕdxdx=xϕ(x)|1/20−∫1/20ϕ(x)⋅1dx+(1−x)ϕ(x)|11/2−∫11/2ϕ(x)⋅(−1)dx=−∫1/20ϕ(x)dx+∫11/2ϕ(x)dx=∫10ϕdudxdx.
Thus, du/dx is the weak derivative of u.
Remark. The original question is that "Compute the weak derivative of the function u of Example 4." However, the weak derivative is given in Example 4 at page 166 so I modify the statement to make it sense. - For x∈Ω=(0,1), function u can be denoted as u(x)=xα. Then u′(x)=αxα−1 for x∈(0,1). Clearly, u∈C1(Ω). Hence it suffices to find suitable α such that ‖u‖21,2<∞. When α>1/2, we find
‖u‖21,2=∫10(|u′|2+|u|2)dx=∫10(α2x2α−2+x2α)dx=α2x2α−12α−1+x2α+12α+1|10=α22α−1+12α+1<∞.
However, when α=1/2, we have‖u‖21,2=∫10(|u′|2+|u|2)dx=∫10(14x−1+x)dx=lnx4+x22|10=∞.
For α<1/2 but α≠−1/2, we have‖u‖21,2=∫10(|u′|2+|u|2)dx=∫10(α2x2α−2+x2α)dx=α2x2α−12α−1+x2α+12α+1|10=∞.
For α=−1/2, we have‖u‖21,2=∫10(|u′|2+|u|2)dx=∫10(14x−3+x−1)dx=−x−28+lnx|10=∞.
Therefore, u∈H1.2(Ω) for and only for α>1/2. - Prove that H1,p0(Rn)=H1,p(Rn) for 1≤p<∞.
- Suppose Ω is a bounded domain, and let S≡C(ˉΩ)⊂X≡Lp(Ω). Pick x0∈Ω. Does the functional Fx0(u)=u(x0) for u∈S extend to a bounded linear functional on X? If not, why not?
- If X is a Hilbert space and S is any subset of X, define
S⊥={y∈X:⟨x,y⟩=0 for all x∈S}.
- Show that S⊥ is a closed subspace of X.
- Show that S∩S⊥ can contain only the zero vector.
- If S⊂T are both subsets of X, show that T⊥⊂S⊥.
- If ˉS is the closure of S in X, show that S⊥=ˉS⊥.
- For y1,y2∈S⊥ and c1,c2∈R, we can find
⟨x,c1y1+c2y2⟩=⟨c1y1+c2y2,x⟩=¯⟨c1y1+c2y2,x⟩=¯c1⟨y1,x⟩+c2⟨y2,x⟩=c1⟨x,y1⟩+c2⟨x,y2⟩for x∈S,
which shows c1y1+c2y2∈S⊥. To show the closeness of S⊥, we let {yn}∞n=1 be a sequence in S⊥ such that {yn}∞n=1 converges to y∈X, i.e., limn→∞‖yn−y‖=0. By the Cauchy-Schwarz inequality, we have|⟨x,y⟩|=|⟨x,y−yn⟩|≤‖x‖‖y−yn‖.
By the squeeze theorem, we have ⟨x,y⟩=0, which means y∈S⊥. Therefore, S⊥ is a closed subspace. - Let x0∈S∩S⊥. Since x0∈S⊥, we know ⟨x,x0⟩=0 for all x∈S. In particular, x0∈S so ⟨x0,x0⟩=0, which implies x0=0 by the definition of inner product. Thus, S∩S⊥⊆{0} and hence S∩S⊥={0}.
- For y∈T⊥, we have ⟨x,y⟩=0 for all x∈T. Since S⊂T, the equality ⟨x,y⟩ holds true for all x∈S. This means y∈S⊥. Therefore, T⊥⊂S⊥.
- Since S⊆ˉS, by (c), we have ˉS⊥⊆S⊥. It suffices to show that S⊥⊆ˉS⊥. Fix y∈S⊥ arbitrarily. For x∈ˉS, there exists a sequence {xn}∞n=1⊆S such that limn→∞‖xn−x‖=0. Note that ⟨xn,y⟩=0 for all n∈N. Then by Cauchy-Scharz inequality, we have
|⟨x,y⟩|=|⟨x−xn,y⟩|≤‖x−xn‖‖y‖.
By the squeeze theorem, we have ⟨x,y⟩=0 for any x∈ˉS, which implies y∈ˉS⊥ and hence S⊥⊆ˉS⊥. The proof is complete. - For vector-valued functions →u=(u1,…,uN) on a domain Ω⊂Rn, it is natrual to define Lp(Ω,RN) as the functions →u:Ω→RN for which |→u|∈Lp(Ω); that is,
‖→u‖p≡(∫Ω|→u|pdx)1/p<∞,where |→u|2=N∑i=1u2i.
Simiarly, if each ui∈C1(Ω), we may define‖∇→u‖p=(∫Ω|∇→u|pdx)1/p,where |∇→u|2=N∑i=1n∑j=1(∂ui∂xj)2.
- Define H1,p0(Ω,RN).
- If N=n, show that div:H1,p0(Ω,Rn)→Lp(Ω) is a continuous linear operator.
- Show that ˜H1,p0(Ω,Rn)≡{→u∈H1,p0(Ω,Rn):div(→u)=0} is a closed subspace of H1,p0(Ω,Rn).
- As for the definition of H1,p0(Rn), H1,p0(Ω,RN) can be defined by
H1,p0(Ω,RN)=¯C10(Ω,RN)
where the bar denotes the completion in the following norm‖→u‖1,p=(‖→u‖pp+‖∇→u‖pp)1/p for →u∈C10(Ω,RN).
- Recall that for →u=(u1,…,un)∈H1,p0(Ω,Rn), div(→u)=n∑i=1∂ui∂xi. Since ∂ui/∂xi∈Lp(Ω), the divergence operator is well-defined. For any →u,→v∈H1,p0(Ω,Rn) and c∈R, we have
div(c→u+→v)=n∑i=1∂∂xi(cui+vi)=cn∑i=1∂ui∂xi+n∑i=1∂vi∂xi=c⋅div(→u)+div(→v),
which means div is a linear operator. By Theorem 1 at page 167 in the textbook, the continuity of a linear operator is equivalent to its boundedness. For nonzero →u∈H1,p0(Ω,Rn) with ‖→u‖1,p=1, the Cauchy–Schwarz inequality implies|n∑i=1∂ui∂xi|p≤np/2(n∑i=1(∂ui∂xi)2)p/2≤np/2|∇→u|p.
Then we obtain‖div(→u)‖p=(∫Ω|n∑i=1∂ui∂xi|pdx)1/p≤(np/2∫Ω|∇→u|pdx)1/p=√n‖∇→u‖p≤√n‖→u‖1,p=√n.
Hence the operator div is bounded. The proof is complete. - Clearly, ˜H1,p0(Ω,Rn) is a subset of H1,p0(Ω,Rn) and is nonempty because the zero function is in ˜H1,p0(Ω.Rn). By (b), it is easy to see that c→u+→v∈˜H1,p0(Ω,Rn) for any →u,→v∈˜H1,p0(Ω,Rn) and c∈R. Thus, ˜H1,p0(Ω,Rn) is a subspace of H1,p0(Ω,Rn). To see ˜H1,p0(Ω,Rn) is closed, we pick →u∈¯˜H1,p0(Ω,Rn). By the definition of closure, there exists a sequence of functions {→uk}∞k=1⊆˜H1,p0(Ω,Rn) such that limk→∞‖→uk−→u‖1,p=0. Since →uk∈˜H1,p0(Ω,Rn), we have div(→uk)=0. By (b), we have
div(→u)=div(limk→∞→uk)=limk→∞div(→uk)=limk→∞0=0.
This implies →u∈˜H1,p0(Ω,Rn). The proof is complete.
Remark. A direct observation is that ˜H1,p0(Ω,Rn)=H1,p0(Ω,Rn)∩div−1({0}). Since div is continuous, the set div−1({0}) is closed, which implies ˜H1,p0(Ω,Rn) is closed. - In Lemma 2 in this section, the vector v∈Y was claimed (but not proved) to be unique. Establish this uniqueness.
- If F:X→Y is a bounded linear operator between normed vector spaces X and Y, show that the nullspace N={x∈X:F(x)=0} is a closed subspace of X.
- If Y is a subspace of a Hilbert space X, define Y⊥ as in Exercise 8. Show (Y⊥)⊥=ˉY.
- Suppose S is a subspace of a Hilbert space X and f:S→R is a linear functional with |f(s)|≤C‖s‖ for all s∈S. Prove that there is a unique linear functional F:X→R extending f (i.e., F(s)=f(s) for s∈S) and preserving the norm:
‖F‖=supx∈X,x≠0|F(x)|‖x‖=sups∈S,s≠0|F(s)|‖s‖.
(Notice that this special case of the Hahn-Banach theorem does not require the axiom of choice.) - Suppose X is a Hilbert space and {xn}∞n=1 is a collection of orthonormal vectors. Given u∈X, define its Fourier coefficients by αn=⟨u,xn⟩.
- Prove Bessel's inequality: ∞∑n=1α2n≤‖u‖2.
- Let Y be the finite-dimensional subspace of X generated by taking linear combinations of x1,…,xN. Show that v=N∑n=1αnxn is the element of Y that minimizes ‖u−v‖ as in Lemma 1 of this section.
- Repeat (b) when Y is infinite dimensional subspace generated by taking linear combinations and limits of all the xn's.
- For any N∈N, we have ‖u−N∑n=1αnxn‖2≥0, which implies
‖u‖2−N∑n=1α2n=‖u‖2−2N∑n=1αn⟨u,xn⟩+N∑n=1α2n≥0.
Here we have used the fact that ⟨xn,xm⟩=0 if n≠m and ⟨xn,xn⟩=1 for all n∈N. Since the sequence N∑n=1α2n is increasing in N and is bounded by ‖u‖2, we obtain Bessel's inequality and complete the proof. - For any v∈Y, there exists c1,…,cN∈R such that v=N∑n=1cnxn. Then we find
‖u−v‖2=‖u‖2−2N∑n=1cn⟨u,xn⟩+N∑n=1c2n=‖u‖2−2N∑n=1αncn+N∑n=1c2n=‖u‖2+N∑n=1[(cn−αn)2−α2n]≥‖u‖2−N∑n=1α2n.
The equality holds when cn=αn for n=1,…,N. Hence N∑n=1αnxn is the unique element of Y that minimizes ‖u−v‖. - It suffices to show that if ∞∑n=1cnxn=u, i.e., limN→∞N∑n=1cnxn=u, then cn=αn for all n∈N. Since the inner product ⟨⋅,⋅⟩ is continuous in both arguments because of the Cauchy-Schwarz inequality. Thus, for any y∈X we have
⟨u,y⟩=⟨∞∑n=1cnxn,y⟩=⟨limN→∞N∑n=1cnxn,y⟩=limN→∞N∑n=1cn⟨xn,y⟩=∞∑n=1cn⟨xn,y⟩.
For y=xm, we get αm=∞∑n=1cn⟨xn,xm⟩=∞∑n=1cnδnm=cm for m∈N. The proof is complete. - A bounded bilinear form on a real Hilbert space X is a map B:X×X→R satisfying
- B(αx+βy,z)=αB(x,z)+βB(y,z),
- B(x,αy+βz)=αB(x,y)+βB(x,z),
- |B(x,y)|≤C‖x‖‖y‖ for all x,y,z∈X and α,β∈R.
- If X is a Hilbert space and T:X→X is a bounded linear operator, the adjoint of T is an operator T∗:X→X defined as follows:
- For y∈X, use the Riesz representation theorem to define T∗y∈X satisfying ⟨Tx,y⟩=⟨x,T∗y⟩ for all x,y∈X.
- Show that T∗:X→X is a bounded linear operator with ‖T∗‖=‖T‖.
- Given y∈X. Then we define the linear functional fy:X→R by fy(x)=⟨Tx,y⟩ for x∈X. Since T is bounded, |fy(x)|≤‖Tx‖‖y‖≤‖T‖‖x‖‖y‖, which implies |fy(x)|/‖x‖≤‖T‖‖y‖. This shows that fy is also bounded. Hence by Theorem 3 (Riesz Representation), there exists a unique element vy such that fy(x)=⟨x,vy⟩ for x∈X. Hence we define T∗:X→X by T∗y=vy for y∈X.
- To get the linearity of T∗, we need to show vcy1+y2=cvy1+vy2 for y1,y2∈X and c∈R, which follows from fcy1+y2=cfy1+fy2. This can be verified by
fcy1+y2(x)=⟨Tx,cy1+y2⟩=c⟨Tx,y1⟩+⟨Tx,y2⟩=cfy1(x)+fy2(x)for x∈X.
This shows that T∗ is a linear operator. Let y∈X be a unit vector. If T∗y=0, then surely ‖T∗y‖≤‖T‖. If T∗y≠0, then we have‖T∗y‖2=⟨T∗y,T∗y⟩=⟨TT∗y,y⟩≤‖TT∗y‖‖y‖=‖TT∗y‖≤‖T‖‖T∗y‖,
which gives ‖T∗y‖≤‖T‖. Taking the supremum over all y with ‖y‖=1, we get ‖T∗‖≤‖T‖. On the other hand, we note that ⟨T∗x,y⟩=⟨y,T∗x⟩=⟨Ty,x⟩=⟨x,Ty⟩, which implies T=(T∗)∗ and hence ‖T‖=‖(T∗)∗‖≤‖T∗‖. Therefore, T∗ is bounded and ‖T∗‖=‖T‖. - If T:X→X is a bounded linear operator on a Hilbert space X, let N={x∈X:T(x)=0} be the nullspace of T and R={Tx:x∈X} be the range of T. Similarly, let N∗ and R∗ denote the nullspace and range of the adjoint T∗ defined in Exercise 16. Prove (N∗)⊥=ˉR (where S⊥ is defined in Exercise 8).
Solution
Solution
Firstly we claim that vw∈Lp′, where p′ satisfies (p′)−1=1−p−1=q−1+r−1. Since v∈Lq(Ω) and w∈Lr(Ω), we have vp′∈Lq/p′(Ω) and wp′∈Lr/p′(Ω). Since 1q/p′+1r/p′=1, we can apply Hölder's inequality to get vp′wp′∈L1(Ω) and‖vw‖p′p′=∫Ω|vw|p′dx=∫Ω|vp′wp′|dx≤‖vp′‖q/p′‖wp′‖r/p′=(∫Ω|v|qdx)p′/q(∫Ω|w|rdx)p′/q=‖v‖p′q‖w‖p′r,
which gives ‖vw‖p′≤‖v‖q‖w‖r. Therefore, we can use Hölder's inequality again to get∫Ω|uvw|dx=∫Ω|u||vw|dx≤‖u‖p‖vw‖p′≤‖u‖p‖v‖q‖w‖r.
The proof is complete.Solution
Since 1p/(λq)+1r/((1−λ)q)=1, we can use Hölder's inequality to get‖u‖q=‖uq‖1/q1=‖uλq⋅u(1−λ)q‖1/q1≤‖uλq‖1/qpλq‖u(1−λ)q‖1/qr(1−λ)q=(∫Ω|uλq|pλqdx)λp(∫Ω|u(1−λ)q|r(1−λ)qdx)(1−λ)r=[(∫Ω|u|pdx)1p]λ[(∫Ω|u|rdx)1r]1−λ=‖u‖λp‖u‖1−λr.
Remark. When Ω has a finite Lebesgue measure, i.e., μ(Ω)<∞, Lr(Ω)⊆Lp(Ω) for p<r. To see this, we let f∈Lr(Ω) and define Ω1={x∈Ω:|f|≤1} and Ω2={x∈Ω:|f|>1}. Then
∫Ω|f|pdμ=∫Ω1|f|pdμ+∫Ω2|f|pdμ≤∫Ω11dμ+∫Ω2|f|rdμ<∞,
which shows f∈Lp(Ω). Here we have used the fact that μ(Ω1)≤μ(Ω) and μ(Ω) is finite. However, it is not true when μ(Ω)=∞ so one may modify the condition for this case.
Solution
Solution
Solution
For u∈C10(Rn), we have ‖u‖1,p<∞, which impliesC10(Rn)⊆{u∈C1(Rn):‖u‖1,p<∞}⊆¯{u∈C1(Rn):‖u‖1,p<∞}=H1,p(Rn).
Taking the clourse in the norm ‖⋅‖1,p, we obtain H1,p0(Rn)=¯C10(Rn)⊆H1,p(Rn). Conversely, let u∈H1,p(Rn), i.e., there exists a sequence {um}∞m=1 such that um∈C1(Rn) with limm→∞‖um−u‖1,p=0. Let ϕ∈C∞0([0,∞)) satisfy ϕ(t)=1 for 0≤t≤1; ϕ(t)=0 for 2≤t<∞ and 0≤ϕ(t)≤1 for t≥0. Then we define um,k(x)=um(x)ϕ(k−1|x|) for x∈Rn and m,k∈N. Clearly, um,k∈C10(Rn). For i∈{1,…,n}, we have∂um,k∂xi(x)=∂um∂xi(x)ϕ(k−1|x|)+um(x)ϕ′(k−1|x|)⋅xik|x|for x∈Rn.
For any fixed m∈N, we find‖∂um,k∂xi−∂um∂xi‖pp=∫Rn|∂um∂xi(x)[ϕ(k−1|x|)−1]+um(x)ϕ′(k−1|x|)⋅xik|x||pdx=∫|x|≥k|∂um∂xi(x)[ϕ(k−1|x|)−1]+um(x)ϕ′(k−1|x|)⋅xik|x||pdx
Recall an elementary inequality: (|a|+|b|)p≤2p−1(|a|p+|b|p) for p≥1. Then we have‖∂um,k∂xi−∂um∂xi‖pp≤2p−1(∫|x|≥k|∂um∂xi|pdx+sup[0,∞)|ϕ′|pkp∫|x|≥k|um|pdx)→0 as k→∞.
Thus, there exists k=k(m)∈N such that um,k(m)∈C10(Rn) satisfies limm→∞‖um,k(m)−u‖1,p=0, i.e., u∈H1,p0(Rn). The proof is complete.Remark. It is plausible to write the range of p based on the definition of H1,p0(Rn) and H1,p(Rn) at the page 166 in the textbook.
Solution
No. Suppose by contradiction that Fx0 can be extended to a bounded linear functional ˜Fx0:X→R. Then for any u∈X with ‖u‖p=1, we have |˜Fx0(u)|≤‖˜Fx0‖, where ‖˜Fx0‖ is a constant independent of u. It is clear that there exists a function ˉu∈C(ˉΩ) satisfying ˉu(x0)=‖˜Fx0‖+1 and ‖ˉu‖p=1. Then we get a contradiction that:‖˜Fx0‖+1=|˜Fx0(ˉu)|≤‖˜Fx0‖.
Therefore, Fx0 cannot be extended to a bounded linear functional on X.Solution
Solution
Solution
Suppose by contradiction that there exist v′∈Y and w′∈X such that u=v′+w′ and ⟨w′,Y⟩=0 but v′≠v. Then from v+w=u=v′+w′, we have w′−w=v−v′∈Y. Since ⟨w′−w,Y⟩=⟨w′,Y⟩−⟨w,Y⟩=0, we have ⟨w′−w,w′−w⟩=0, which means w′=w and hence v′=v. This leads a contradiction and the uniqueness of v and w follows.Solution
For x1,x2∈N and c∈R, we have F(cx1+x2)=cF(x1)+F(x2)=c⋅0+0=0, which implies cx1+x2∈N. Moreover, since N⊆X and 0∈N, N is a subspace of X. Note that F is continuous linear operator due to its boundedness by Theorem 1 at page 167 in the textbook. Since {0} is a closed set, we have N=F−1({0}) is also closed. Therefore, N is a closed subspace and we complete the proof.Remark. Let {xn}∞n=1 be a sequence in N and converges to x∈X, i.e., F(xn)=0 for all n∈N. Then by the continuity of F, we have F(x)=F(limn→∞xn)=limn→∞F(xn)=limn→∞0=0, which implies x∈N, i.e., N is a closed set. The proof is complete.
Solution
Recall that the closure of a set A is the smallest closed set containing A, i.e., if A⊆B and B is closed, then ˉA⊆B. For y∈Y, we know ⟨y,x⟩=0 for all x∈Y⊥. Moreover, we have⟨x,y⟩=¯⟨y,x⟩=0 for any x∈Y⊥.
This means y∈(Y⊥)⊥, i.e., Y⊆(Y⊥)⊥. Hence we have ˉY⊆(Y⊥)⊥. It suffices to show that (Y⊥)⊥⊆ˉY. For any y∈(Y⊥)⊥, then we have ⟨x,y⟩=0 for all x∈Y⊥, which implies y∈ˉY, i.e., (Y⊥)⊥⊆ˉY. The proof is complete.Solution
By the Riesz Representation theorem (Theorem 3), there exists a unique element xf∈S such thatf(s)=⟨s,xf⟩for all s∈S.
Then we define the extension of f to X byF(x)=⟨x,xf⟩for x∈X.
Clearly, F is linear functional and F(s)=f(s) for s∈S. Moreover, the norm of linear operator satisfies‖F‖=‖xf‖=‖f‖=sups∈S,s≠0|f(s)|‖s‖=sups∈S,s≠0|F(s)|‖s‖.
Suppose there ia an another linear extension F′ of f to X. Then by the Riesz Representation Theorem again, there exists an element xF′∈X such that F′(x)=⟨x,xF′⟩ for all x∈X. Since F′(s)=f(s)=F(s) for all s∈S, we have ⟨s,xf−xF′⟩=0 for all s∈S and hence xf−xF′∈S⊥. But we find‖xf‖=‖F‖=‖F′‖=‖xF′‖=‖xf+(xF′−xf)‖=√‖xf‖2+‖xF′−xf‖2,
which shows ‖xF′−xf‖=0 and hence xF′=xf and F=F′.Solution
Solution
Fix x∈X arbitrarily. Then we define Fx:X→R by Fx(y)=B(x,y) for y∈X. By (ii), we knowFx(cy1+y2)=B(x,cy1+y2)=cB(x,y1)+B(x,y2)=cFx(y1)+Fx(y2),
which shows Fx is a linear operator on X. In addition, by (iii), we know ‖Fx‖≤C‖x‖, which means Fx is bounded. Hence by Theorem 3 (Riesz Representation), there exists a unique vx∈X such that Fx(y)=⟨y,vx⟩ for all y∈X. Note that ⟨y,vx⟩=⟨vx,y⟩ because X is a real Hilbert space. Now we define the operator A:X→X by Ax=vx for all x∈X. It suffices to show that A is a bounded linear operator. For x1,x2∈X and d∈R, it remains to show vdx1+x2=dvx1+vx2, which is equivalent to verify that Fdx1+x2=dFx1+Fx2. This follows directly from (i):Fdx1+c2(y)=B(dx1+x2,y)=dB(x1,y)+B(x2,y)=dFx1(y)+Fx2(y).
Hence A is a linear operator. By (iii), we have|⟨Ax,y⟩|=|B(x,y)|≤C‖x‖‖y‖,
which implies ‖Ax‖2≤C‖x‖‖Ax‖ and ‖Ax‖/‖x‖≤C. This means A is bounded. To end this proof, we need to establish the uniqueness of A. Suppose by contradicion that there exists an another bounded linear operator A′:X→X such that B(x,y)=⟨A′x,y⟩ for all x,y∈X. Then we have ⟨(A′−A)x,y⟩=0 for all x,y∈X. Taking y=(A′−A)x, it gives ‖(A′−A)x‖=0 and hence A′x=Ax for all x∈X, i.e., A′=A. This leads a contradiction. Therefore, we complete the proof.Remark. The linearity of A can be proved as follows. For x1,x2∈X and c∈R, we have
⟨A(cx1+x2),y⟩=B(cx1+x2,y)=cB(x1,y)+B(x2,y)=c⟨Ax1,y⟩+⟨Ax2,y⟩=⟨cAx1+Ax2,y⟩,
which gives ⟨A(cx1+x2)−(cAx1+Ax2),y⟩=0 for all y∈X. Thus, A(cx1+x2)=cAx1+Ax2.Solution
Solution
We firstly prove that R⊆(N∗)⊥. Let y∈R. Then by the definition of the range, there is an element x∈X such that y=Tx. For any z∈N∗, we find⟨z,y⟩=⟨y,z⟩=⟨Tx,z⟩=⟨x,T∗z⟩=⟨x,0⟩=0.
This means y∈(N∗)⊥ and R⊆(N∗)⊥. By the part (a) of Exercise 8, ˉR⊆(N∗)⊥. Here we have used the fact that if A⊆B and B is closed, then ˉA⊆B. On the other hand, we want to prove that (N∗)⊥⊆ˉR, which is equivalent to show that (ˉR)c⊆((N∗)⊥)c. Here Ac denotes the complement set of A. Suppose by contradiction that y∈(ˉR)c but y∈(N∗)⊥. By Lemma 2 (The Projection Theorem), there exists a unique vy∈ˉR and wy∈X such that y=vy+wy and ⟨wy,ˉR⟩=0, where wy≠0. The condition ⟨wy,ˉR⟩=0 implies ⟨wy,Tu⟩=0 and hence ⟨T∗wy,u⟩=0 for all u∈X. Thus, wy∈N∗. From y∈(N∗)⊥, we have0=⟨z,y⟩=⟨z,vy+wy⟩=⟨z,vy⟩+⟨z,wy⟩=⟨z,wy⟩for z∈N∗,
which means wy∈(N∗)⊥. Here we have used the facts that vy∈ˉR and there is an element uy∈X such that Tuy=vy and⟨z,vy⟩=⟨vy,z⟩=⟨Tuy,z⟩=⟨uy,T∗z⟩=⟨uy,0⟩=0.
Hence we arrive at wy∈N∗∩(N∗)⊥. This means wy=0 and leads a contradiction. Therefore, we complete the proof of (N∗)⊥=ˉR.
沒有留言:
張貼留言