Tải bản đầy đủ (.pdf) (185 trang)

One dimensional ergodic schrödinger operators i general theory(pg 31 215)

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (1.31 MB, 185 trang )

<span class="text_page_counter">Trang 1</span><div class="page_container" data-page="1">

(a) Prove Hăolders inequality: If <sup>1</sup><i><sub>p</sub></i> + <sup>1</sup><i><sub>q</sub>= 1, φ∈ <small>p</small></i>(<i>Z), and ψ ∈ <small>q</small></i>(Z),

<i>then φψ∈ </i><small>1</small>(Z) and

<i>Hints: If either φ or ψ vanishes identically, the claim is trivial.</i>

Otherwise, reduce to the case of unit vectors and apply Young’s inequality.

(b) Prove that<i> · pis a norm. Hint : First show thatφp</i>= sup<i>φ, ψwhere ψ ranges over all unit vectors in <small>q</small></i>and<i>φ, ψ =</i><sup></sup><i>φ(n)ψ(n).(c) Prove that (<sup>p</sup></i>(<i>Z),  · p</i>) is a Banach space.

<i><b>Exercise 1.2.10. Given Hausdorff topological spaces X and Y , let C(X, Y )</b></i>

<i>denote the space of continuous functions from X to Y , C(X) = C(X,</i>C) the

<i>space of continuous complex-valued functions on X, and Cc(X) the set ofall f∈ C(X) such that supp(f) := {x : f(x) = 0} is compact.</i>

<i>(a) If X is compact, show that</i>

<i>defines a norm on C(X) that makes C(X) into a Banach space.(b) If X is not compact, show that Cc(X) is not complete. Identify</i>

<i>the completion of C<sub>c</sub>(X) as an explicit subspace of C(X). Hint :Consider the one-point compactification X := X∪ {∞}.</i>

<b>Exercise 1.2.11. Show that every Hilbert space has a maximal orthonormal</b>

subset and that a Hilbert space is separable if and only if every orthonormal subset of <i>H is countable. Hint: Zorn.</i>

defines an inner product with respect to which<i>H is a Hilbert space.</i>

(b) Give necessary and sufficient conditions on Γ and (Hγ)γ<i><sub>∈Γ</sub></i> for <i>H</i>

</div><span class="text_page_counter">Trang 2</span><div class="page_container" data-page="2">

<i>1.3. Bounded Linear Operators</i> 15

in the sense that there is a linear bijection between them that pre-serves inner products.

<i><small>γ∈Γ</small>Hγthe (external) direct sum (or orthog-onal sum). If Γ ={1, 2, . . . , n} for some n ∈ N, we write the direct sum asH</i><small>1</small><i>⊕ · · · ⊕ Hn.</i>

<b>Notes and Remarks. Banach spaces received their name in the </b>

mono-graph of Fr´<b>echet [96], in honor of foundational work of Banach on normedlinear spaces [19]. Hilbert spaces are named for fundamental contributionsof Hilbert to spectral theory, such as the proof of the spectral theorem [122].</b>

<i><b>Schoenflies [221] used the term “Hilbertscher Raum” to refer to the unit ball</b></i>

<i>of </i><sup>2</sup>(N).

<b>The name of Theorem 1.2.4 comes from contributions of Cauchy [53],who proved the finite-dimensional version, and Schwarz [224], who proved it</b>

<i>for functions in C([0, 1]). It turns out that Bouniakowsky had the inequality</i>

<i><b>on C([0, 1]) before Schwarz [43]. Bessel’s inequality (formula (1.2.24) from</b></i>

<b>Theorem 1.2.14) gets its name from Bessel’s work [34] on trigonometric</b>

<b>The proof of Proposition 1.2.9 is an argument due to F. Riesz [205] andRellich [200]. Their arguments discuss complementation of closed subspaces</b>

in a Hilbert space, but the changes needed to discuss convex sets are merely cosmetic.

<i>Completeness of L<sup>p</sup></i> is proved in most graduate-level real analysis books,

<b>so we omit the proof here; see, e.g., Royden [210], Rudin [211], Simon [244],</b>

<i><b>or Stein and Shakarchi [250]. Even though completeness of </b></i><sup>2</sup> is a corollary of this statement (by considering a space equipped with the counting mea-sure), we elected to present the proof, since it is central to the text, whereas

<i>completeness of general L<sup>p</sup></i> spaces does not figure as prominently here.

<b>1.3. Bounded Linear Operators</b>

We now introduce linear operators on normed spaces. The primary goal of this section is to set notation and review background results to set the stage for the spectral theorem.

from <i>V to W is a pair (D(A), A), where D(A) is a subspace of V, calledthe domain of A, and A is a linear map from D(A) toW, that is, we haveA(aφ + bψ) = aA(φ) + bA(ψ) for φ, ψ∈ D(A) and a, b ∈ C. For linearoperators, we will generally drop the parentheses and write Aφ for A(φ).We will often just write A for the linear operator (D(A), A), especially if</i>

the domain is clear from context.

</div><span class="text_page_counter">Trang 3</span><div class="page_container" data-page="3">

<i>The range of A is</i>

(1.3.1) <i>Ran(A) ={Aφ : φ ∈ D(A)},and the kernel of A is</i>

(1.3.2) <i>Ker(A) ={φ ∈ D(A) : Aφ = 0}.A linear operator is called bounded if D(A) =V and</i>

(1.3.3) <i>A := sup{Aφ : φ ∈ V, φ = 1} < ∞.</i>

<i>The quantity in (1.3.3) is called the operator norm of A. The space of</i>

all bounded operators from <i>V to W is denoted by B(V, W), in the event</i>

that <i>V = W, we say that A is a bounded operator in V and we writeB(V) := B(V, V).</i>

The space <i>B(V, W) inherits natural vector space operations. To wit:given A, B∈ B(V, W) and a ∈ C, one can define A + B, aA ∈ B(V, W) via</i>

(1.3.4) <i>(A + B)v = Av + Bv,(aA)v = a(Av).</i>

In addition to vector space operations, composition of functions endows spaces of bounded operators with a multiplicative structure. Concretely, if <i>V, W, and X are normed spaces, A ∈ B(V, W), and B ∈ B(W, X ),then BA∈ B(V, X ) (where (BA)v = B(Av)); indeed, since BAv ≤BAv ≤ BAv for all v ∈ V, one has</i>

<i>We denote by I<small>V</small>the identity operator I<small>V</small>ψ = ψ (and drop the subscriptwhen the underlying space is clear). An operator A∈ B(V, W) is calledinvertible if there is an element B∈ B(W, V) such that AB = I<sub>W</sub></i> and

<i>BA = I<small>V</small><sup>; we write B = A</sup><sup>−1</sup><sup>and call B the inverse of A.</sup></i>

<i>collection of m× n matrices with complex entries.</i>

It is often of interest to extend a linear operator from a subspace to the whole space. If the operator is bounded on the subspace and the subspace is dense, there is always a unique bounded extension.

<i>(D(A), A) is a linear operator inV such that D(A) is dense in V and</i>

(1.3.6) <i>AD(A)</i>:= sup<i>{Ax : x ∈ D(A) and x = 1} < ∞.</i>

<i>There is a unique A∈ B(V) such that Ax = Ax for all x∈ D(A). Moreover, A = AD(A).</i>

</div><span class="text_page_counter">Trang 4</span><div class="page_container" data-page="4">

<i>1.3. Bounded Linear Operators</i> 17

<i>D(A) such that ϕ<sub>n</sub>→ ψ. Since</i>

<i>Aϕn− Aϕm = A(ϕn− ϕm</i>)<i> ≤ AD(A)ϕn− ϕm,η = lim Aϕnexists by completeness. Moreover, if (ϕ<small></small></i>

<i><small>n=1</small></i> is another

<i>se-quence converging to ψ, the calculationAϕn− Aϕ<small>n = A(ϕn</small>− ϕ<small></small></i>

<i><small>n</small></i>)<i> ≤ AD(A)ϕn− ϕ<small>n</small>implies that Aϕ<small></small></i>

<i><small>n</small>→ η as well, so the limit is independent of the choice ofthe sequence converging to ψ. Thus, we define Aψ = η. One can check that</i>

<i>A∈ B(V, W) in norm if An− A → 0 as n → ∞. We say that An→ Astrongly, denoted</i>

<i>A = s-lim<small>n→∞</small><sup>An,</sup></i>

<i>if Anψ→ Aψ for every ψ, that is (A − An)ψ → 0 as n → ∞ for every ψ.</i>

The proof of the following proposition is left to the reader.

(a) <i>B(V, W) is a vector space and (1.3.3) defines a norm on B(V, W)satisfying (1.3.5).</i>

<i>(b) IfW is a Banach space, so too is B(V, W).</i>

<i>(c) IfW is a Banach space and A</i><small>1</small><i>, A</i><sub>2</sub><i>, . . . is a sequence of elements ofB(V, W) that is absolutely summable in the sense that</i>

</div><span class="text_page_counter">Trang 5</span><div class="page_container" data-page="5">

<b>Proof. Exercise 1.3.2.</b>  Throughout the remainder of this section,<i>V denotes a Banach space. Abounded operator is continuous, that is, if φ<sub>n</sub>→ φ, then Aφn→ Aφ. As a</i>

sample application of Proposition 1.3.5, we have the following.

<i>ofB(V). If A ∈ G (V) and B < A<small>−1</small><small>−1</small><sub>, then A + B is invertible. In</sub>particularG (V) is an open subset of B(V).</i>

<i>I + A<small>−1</small><sub>B. By assumption</sub>A<small>−1</small><sub>B</sub> ≤ A<small>−1</small>B < 1, so one can check that</i>

<i>converges to the inverse of I + A<small>−1</small><sub>B with the help of Proposition 1.3.5.</sub></i> 

<i>complex numbers z for which A− zI is invertible. The spectrum of A is</i>

(1.3.11) <i>σ(A) =C \ ρ(A) = {z ∈ C : A − zI is not invertible}.</i>

<i>A scalar z∈ C is called an eigenvalue of A if Av = zv for some nonzerov∈ V; equivalently, z is an eigenvalue of A if and only if Ker(A − zI) isnontrivial. Let us note that every eigenvalue of A necessarily belongs toσ(A), but the converse need not hold in infinite-dimensional spaces, as we</i>

will see later.

<i>The discrete spectrum of A consists of all isolated eigenvalues of A havingfinite multiplicity. That is, z∈ σ</i><small>disc</small><i>(A) if and only if</i>

(1.3.12) <i>0 < dim Ker(A− zI) < ∞</i>

<i>and B(z, δ)∩ σ(A) = {z} for some δ > 0 (recall that B(z, δ) denotes theopen ball of radius δ centered at z). The complement of σ</i><sub>disc</sub><i>(A) in σ(A) iscalled the essential spectrum of A and is denoted σ</i><sub>ess</sub><i>(A) = σ(A)\ σ</i><small>disc</small><i>(A).The spectral radius of A is given by</i>

(1.3.13) <i>spr(A) = sup{|z| : z ∈ σ(A)}.</i>

<i>Later on, we will see that σ(A) is compact, so sup can be replaced by max</i>

in the definition of the spectral radius.

<i>The resolvent of A is the following map</i>

(1.3.14) <i>R(A,·) : ρ(A) → B(V), z → (A − z)<sup>−1</sup>.</i>

</div><span class="text_page_counter">Trang 6</span><div class="page_container" data-page="6">

<i>1.3. Bounded Linear Operators</i> 19

Here and throughout the text, we adopt a standard convention from the mathematical physics literature and suppress occurrences of the

<i>iden-tity operator, e.g., writing (A− z)<small>−1</small><sub>instead of (A</sub>− zI)<small>−1</small></i><sub>. The following</sub>

elementary identities will be very useful.

<b>Proposition 1.3.8 (Resolvent identities).</b>

<i>(a) If A∈ B(V) and z, z<small></small>∈ ρ(A), then</i>

<i>R(A, z)− R(A, z<sup></sup>) = (z− z<sup></sup>)R(A, z)R(A, z<small></small></i><sub>)</sub> <i>= (z− z<small></small><sub>)R(A, z</sub><small></small><sub>)R(A, z).</sub></i>

<i>(b) If A, B∈ B(V) and z ∈ ρ(A) ∩ ρ(B), then</i>

<i>R(A, z)− R(B, z) = R(A, z)(B − A)R(B, z)= R(B, z)(B− A)R(A, z).</i>

Let us note the following pleasant consequence of Proposition 1.3.8. The

<i>function z→ R(A, z) defines an analytic map from the open set ρ(A) ⊂ C</i>

to<i>B(H). Concretely, (1.3.15) implies thatd</i>

<i>dz<sup>R(A, z) := lim</sup><small>w→z</small></i>

<i>R(A, w)− R(A, z)</i>

<i>w− z<sup>= R(A, z)</sup></i><sup>2</sup>

<i>for any z∈ ρ(A). Many facts about analytic functions carry over from the</i>

scalar case to the case of Banach-space-valued functions; we make use of this in the following proofs.

<i>and σ(A) is a nonempty compact subset ofC.</i>

<i>One can check that R∈ B(V) and R(A − z) = (A − z)R = I, whencez∈ ρ(A). Thus, σ(A) is bounded and (1.3.17) holds. By Theorem 1.3.6,ρ(A) is open. Thus, σ(A) is compact.</i>

<i>It remains to see that σ(A) is nonempty. To that end, suppose to thecontrary that σ(A) =∅ and define g(z) = (A − z)<small>−1</small><sub>for each z</sub>∈ ρ(A) = C.</i>

By (1.3.18), one has <i>g(z) = O(1/|z|) as |z| → ∞. Thus, by Liouville’stheorem (Exercise 1.3.15), g(z) = 0 for all z∈ C, a contradiction.</i> 

</div><span class="text_page_counter">Trang 7</span><div class="page_container" data-page="7">

<i>If A∈ B(V) and p(z) = cnz<small>n</small></i>+<i>· · ·+c</i><small>1</small><i>z +c</i><small>0</small> is a polynomial, it is natural to define

(1.3.19) <i>p(A) = c<sub>n</sub>A<sup>n</sup></i>+<i>· · · + c</i><small>1</small><i>A + c</i><sub>0</sub><i>I.</i>

<i>It is helpful to note that p(A) and q(A) commute with one another for allpolynomials p and q and that the spectrum of A transforms in a natural</i>

fashion under this operation.

<i>p is a polynomial, then</i>

(1.3.20) <i>σ(p(A)) = p(σ(A)) :={p(z) : z ∈ σ(A)}.</i>

<i>σ(p(A)) = σ(c</i><small>0</small><i>I) ={c</i><small>0</small><i>} = p(σ(A)).</i>

<i>Henceforth, assume deg(p)≥ 1. First, suppose w ∈ p(σ(A)), so w = p(E)with E∈ σ(A). One can factor</i>

<i>p(z)− w = (z − E)q(z)</i>

<i>for a nonzero polynomial q. Since A− E is not invertible, neither isp(A)− p(E) = (A − E)q(A) = q(A)(A − E),</i>

<i>proving w = p(E)∈ σ(p(A)) and thus p(σ(A)) ⊆ σ(p(A)).Conversely, let w∈ C \ p(σ(A)), and write</i>

<i>for each n, proving</i>

<i><small>n→∞</small>A<small>n</small><small>1/n</small>.</i>

</div><span class="text_page_counter">Trang 8</span><div class="page_container" data-page="8">

<i>1.3. Bounded Linear Operators</i> 21

Then, it remains to show

<i><small>n→∞</small>A<small>n</small><small>1/n</small>≤ spr(A).</i>

<i>To that end, consider the function f (z) = (I− zA)<small>−1</small></i><sub>, which is analytic on</sub> <i>D ={z ∈ C : z spr(A) < 1}, and which enjoys the series representation</i>

<i>thereupon. Denoting the radius of convergence of this series by R, Cauchy’sformula for R (Exercise 1.3.15) yields</i>

lim sup

<i><small>n→∞</small>A<small>n</small><small>1/n</small></i> = <sup>1</sup>

operations are continuous in <i>B = B(V, W). That is, if an→ a ∈ C,A<sub>n</sub>→ A ∈ B, and Bn→ B ∈ B, then An+ B<sub>n</sub>→ A + B, AnB<sub>n</sub>→ AB,and a<sub>n</sub>A<sub>n</sub>→ aA.</i>

<b>Exercise 1.3.2. Prove Proposition 1.3.5.Exercise 1.3.3. Prove Proposition 1.3.8.</b>

<i><small>n=1</small></i>is a sequence of bounded linear operators in a Banach space <i>V. If supnAn < ∞, show that (An)n<small>≥1</small></i> converges strongly if there is a dense subset <i>V</i><small>0</small><i>⊆ V such that Anψ is convergent inVfor all ψ∈ V</i><small>0</small>.

<i><small>n=1</small></i> be a bounded sequence of complex

<i>num-bers, and define a linear operator D : </i><sup>2</sup>(<i>N) → </i><small>2</small>(<i>N) by [Dψ](n) = dnψ(n).Characterize σ(D) in terms of the sequence d.</i>

</div><span class="text_page_counter">Trang 9</span><div class="page_container" data-page="9">

<i><b>Exercise 1.3.7. Let (Ω,</b>B, μ) be a σ-finite measure space, and let g ∈L<small>∞</small><sub>(μ). Define a linear operator M = M</sub></i>

<i><small>g</small>: L</i><sup>2</sup><i>(μ)→ L</i><small>2</small><i>(μ) by [M f ](ω) =g(ω)f (ω). Show that M is a bounded operator. Characterize σ(M ) in termsof g.</i>

<i>I− AB is invertible if and only if I − BA is invertible. Hint: Looking at</i>

the case when <i>AB < 1 might be helpful (though it is not necessary).</i>

<i>(a) Show that Ker(A) and Ran(A) are linear subspaces ofV and W,</i>

(b) True or false: If <i>V and W are Banach spaces and A is bounded,then Ker(A) and Ran(A) are closed subspaces. (Give a proof or</i>

counterexample in each case.)

such that the following hold.

<i>(a) A has an isolated eigenvalue of infinite multiplicity.</i>

<i>(b) A has eigenvalues of finite multiplicity that are dense in [0, 1].(c) A has eigenvalues of infinite multiplicity that are dense in [0, 1].</i>

Show that

<i>dist(z, σ(A))<sup>.</sup></i>

<i>Recall that the distance from a point to a set is given by dist(z, σ(A)) :=</i>

inf<i>{|z − s| : s ∈ σ(A)}. Hint: Theorem 1.3.6.</i>

<i>γ∈ Γ, let H =</i><sup></sup><i><sub>γ</sub><sub>∈Γ</sub>Hγ</i> denote their direct sum (cf. Definition 1.2.15), and

<i>suppose we are given bounded operators Aγ∈ B(Hγ) for each γ∈ Γ suchCharacterize σ(A) in terms of σ(A<sub>γ</sub></i>).

<b>Definition 1.3.12. We denote the operator constructed in Exercise 1.3.12</b>

by 

<i><small>γ∈Γ</small><sup>Aγ</sup><sup>and call it the direct sum of the Aγ.</sup></i>

</div><span class="text_page_counter">Trang 10</span><div class="page_container" data-page="10">

<i>1.3. Bounded Linear Operators</i> 23

real vector space <i>V, and suppose p : V → R<sub>≥0</sub></i> is positive homogeneous and

<i>subadditive. Recall that p is positive homogeneous if and only ifp(av) = ap(v) for all a≥ 0 and v ∈ V</i>

and is subadditive if and only if

<i>p(v + w)≤ p(v) + p(w) for all v, w ∈ V.</i>

<i>Suppose </i><sub>0</sub> : <i>U → R is a linear functional such that </i><small>0</small><i>(u)≤ p(u) for allu∈ U. Show that </i><small>0</small> enjoys an extension to all of<i>V that is also bounded byp in two steps:</i>

<i>(a) If w∈ V \ U, let W denote the subspace generated by w and U.Show that there exists  :W → R linear such that |<sub>U</sub>= </i><sub>0</sub> and

<i>(x)≤ p(x) for all x ∈ W. Hints: The extension is uniquely definedby (w). Show that one can choose (w) in such a manner that</i>

<i>(u</i><small>1</small><i>+ w)≤ p(u</i><small>1</small><i>+ w) for all u</i><small>1</small><i>∈ U,</i>

<i>(u</i><small>2</small><i>− w) ≤ p(u</i><small>2</small><i>− w) for all u</i><small>2</small><i>∈ U,</i>

<i>and that this defines the desired .</i>

<i>(b) Show that there exists an extension of </i><sub>0</sub> to all of <i>V that is alsobounded by p. Hint : Zorn.</i>

of a complex vector space <i>V, and suppose p : V → R≥0</i> is subadditive and

<i>satisfies p(av) =|a|p(v) for all a ∈ C and v ∈ V. Suppose </i><small>0</small> :<i>U → C is a</i>

linear functional such that <i>|</i><small>0</small><i>(u)| ≤ p(u) for all u ∈ U. Show that </i><small>0</small> enjoys an extension to all of<i>V that is also bounded by p. Hint: Write </i><small>0</small> <i>= </i><sub>1</sub><i>+ i</i><sub>2</sub> <i>for real linear functionals j, and show that </i><small>2</small> <i>uniquely determines </i><small>1</small> via (1.3.34) <i></i><small>2</small><i>(v) =−</i><small>1</small><i>(iv),v∈ V.</i>

<i>set. A function f : D→ V is said to be continuously differentiable on D if</i>

<i>f (w)− f(z)w− z</i>

<i>exists for all z∈ D and defines a continuous map D → V.</i>

(a) Formulate and prove a Banach-space-valued version of Liouville’s

<i>theorem. Hint : If f :C → V is bounded and continuously differen-tiable, the same is true of ◦ f for any  ∈ B(V, C).</i>

(b) Formulate and prove a Banach-space-valued version of the Cauchy

<i>integral formula. Hint : Reduce to</i>C via linear functionals as in the previous part.

</div><span class="text_page_counter">Trang 11</span><div class="page_container" data-page="11">

(c) Formulate and prove a Banach-space-valued version of Cauchy’s

<i>formula for the radius of convergence of a power series. Hint : </i>

Start-ing from the Cauchy integral formula, this should look fairly similar to the standard proof from complex analysis.

<b>Notes and Remarks. Theorem 1.3.11 is due to Gelfand [100]. The </b>

ex-tension result in Exercise 1.3.13 is a version of the Hahn–Banach theorem,

<b>named for work of Hahn in 1927 [113] and Banach in 1929 [20, 21]. The</b>

extension to complex functionals in Exercise 1.3.14 is generally credited to

<b>Bohnenblust and Sobczyk [39], Murray [179], Sukhomlinov [255]; the key</b>

observation (1.3.34) appeared earlier in work of Lă<b>owig [170].1.4. Self-Adjoint Operators and the Spectral Theorem</b>

We now arrive at the fundamental result of the first portion of the chapter: the spectral theorem for bounded self-adjoint operators. In fact, we will see several theorems, each of which might be considered a facet of the spectral theorem. Each version of the spectral theorem will be used in an essential manner throughout the text. The multiplicity of results and perspectives will, in sum, comprise our understanding of “the spectral theorem”.

We begin by defining fundamental objects, namely the dual space and adjoint of a bounded operator.

<i>dual space ofV; elements of V<small>∗</small><sub>are called bounded linear functionals on</sub>V. In view of Proposition 1.3.5, V<small>∗</small></i> <sub>is a Banach space with respect to the</sub>

operator norm.

Throughout the remainder of this section,<i>H denotes a separable Hilbertspace. For each φ∈ H,</i>

defines an element of <i>H<small>∗</small></i><sub>. In fact, every element of</sub><i>H<small>∗</small></i> <sub>is of this form.</sub>

<i> = <sub>φ</sub>. The map T : φ→ φis an isometric anti-linear</i><small>1</small> <i>bijection fromHontoH<small>∗</small><sub>.</sub></i>

subspace of <i>H. Choose a unit vector χ ∈ [Ker ]<small>⊥</small><sub>and take φ = (χ)χ. We</sub>have  = φand φ is the unique element ofH with this property </i>

(Exer-cise 1.4.1). Isometry and anti-linearity follow from the properties defining an inner product, which we leave to the reader. 

<small>1</small><i><small>By “anti-linear”, we mean T (φ + ψ) = T φ + T ψ and T (aφ) = aT φ for all φ, ψ∈ H and alla∈ C.</small></i>

</div><span class="text_page_counter">Trang 12</span><div class="page_container" data-page="12">

<i>1.4. Self-Adjoint Operators and the Spectral Theorem</i> 25

there is a unique element <i>ψ∈ H such that</i>

<i>for all φ∈ H. The map A<small>∗</small><sub>: ψ</sub>→ ψ is called the adjoint of A.</i>

<i>B(C<small>n</small></i>) = C<i><small>n×n</small></i><sub>. In this case, one can check that the adjoint operation</sub> <i>coincides with the conjugate transpose. That is, (A<small>∗</small></i><sub>)</sub>

<i><small>jk</small>= A<sub>kj</sub>for A∈ C<small>n×n</small></i><sub>.</sub>

Let us collect a few of the main properties of the adjoint operation. The following proposition is left to the reader.

<i>unitary if AA<small>∗</small><sub>= A</sub><small>∗</small><sub>A = I, normal if AA</sub><small>∗</small><sub>= A</sub><small>∗</small><sub>A, and idempotent if</sub>A</i><sup>2</sup><i>= A.</i>

<i>self-adjoint, then σ(A) is a compact subset ofR. Moreover, for any normalA∈ B(H),</i>

<i><b>Proof. Suppose A is self-adjoint. In view of Proposition 1.3.9, we need only</b></i>

<i>show that σ(A)⊆ R to prove the first statement. Given z = x + iy withy = 0, notice that, for any ψ, one has</i>

(1.4.7) <i>(A − z)ψ</i><small>2</small> =<i>(A − x)ψ</i><small>2</small>+<i>|y|</i><small>2</small><i>ψ</i><small>2</small><i>≥ |y|</i><small>2</small><i>ψ</i><small>2</small><i>.</i>

<i>In particular, this implies that (A− z) is one-to-one. Since Im ¯z = 0also, this implies Ker(A− ¯z) = {0} as well.</i>

<i>Additionally, (1.4.7) implies that Ran(A− z) is closed. For, if φn∈Ran(A− z) and φn→ φ, writing φn= (A− z)ψn</i> and making use of (1.4.7) gives us

<i>ψn− ψm</i><small>2</small><i>≤ |y|<sup>−2</sup>(A − z)(ψn− ψm)</i><small>2</small>=<i>|y|<sup>−2</sup>φn− φm</i><small>2</small><i>,</i>

</div><span class="text_page_counter">Trang 13</span><div class="page_container" data-page="13">

<i>showing that ψn→ ψ for some ψ ∈ H, and hence φ = (A − z)ψ belongs toRan(A− z).</i>

<i>Since Ran(A− z) is closed, we may employ Theorem 1.2.11 to arrive atRan(A− z) = Ran(A − z) =</i><sup></sup><i>Ran(A− z)</i><sup></sup><i><sup>⊥⊥</sup></i>

<i>Ker(A<small>∗</small>− ¯z)</i><sup></sup><i><sup>⊥</sup></i>=<i>{0}<sup>⊥</sup></i>=<i>H.</i>

<i>Thus, A− z is bijective. Finally, note that (1.4.7) implies (A − z)<small>−1</small></i> <sub>is</sub>

bounded: <i>R(A, z) ≤ |y|<small>−1</small><sub>. Thus, z</sub>∈ ρ(A), as desired.</i><small>2</small>

<i>Now, if A is normal, using (1.4.4) gives us</i>

<i>A</i><small>2</small><i> = (A<sup>∗</sup></i>)<sup>2</sup><i>A</i><sup>2</sup><i><small>1/2</small></i>=<i>(A<sup>∗</sup>A)</i><sup>2</sup><i><small>1/2</small></i> =<i>A<sup>∗</sup>A = A</i><small>2</small><i>.</i>

Inductively, <i>A</i><small>2</small><i><small>n</small></i>

<i> = A</i><small>2</small><i><small>n</small></i>

<i>for all n∈ Z</i><small>+</small>. Thus, <i>A = spr(A) follows</i>

Orthogonal projections (see Definition 1.2.10) play an important role in the spectral theorem. Let us collect a few important properties here.

<i>orthogonal projection ontoV.</i>

<i>where we explicitly allow the case in which both sides are∞. Inparticular, the sum does not depend on the choice of basis.</i>

is precisely the orthgonal projection onto the closed subspace <i>V<small>⊥</small></i><sub>). For any</sub> <i>ψ, since Qψ∈ V<small>⊥</small></i><sub>, we have</sub>

<i>ψ</i><small>2</small> =<i>P ψ</i><small>2</small>+<i>Qψ</i><small>2</small> <i>≥ P ψ</i><small>2</small><i>,</i>

showing<i>P  ≤ 1, and hence P ∈ B(H), since P is readily seen to be linear,</i>

which proves (a).

<i>For any φ∈ V, P φ = φ by definition. In particular, given ψ ∈ H,P ψ∈ V, so the previous observation implies P</i><small>2</small><i>ψ = P ψ, showing P</i><small>2</small> <i>= P .Next, using P ψ⊥ Qφ and P φ ⊥ Qψ, one notes</i>

<i>φ, P ψ = P φ + Qφ, P ψ = P φ, P ψ = P φ, P ψ + Qψ = P φ, ψ,showing P = P<small>∗</small></i> <sub>and concluding the proof of (b).</sub>

<small>2By the closed graph theorem, one does not actually need to check boundedness of the inversehere. We elected to directly check it since we have not stated or proved the closed graph theoremin the text.</small>

</div><span class="text_page_counter">Trang 14</span><div class="page_container" data-page="14">

<i>1.4. Self-Adjoint Operators and the Spectral Theorem</i> 27

To prove (1.4.8), first note that <i>φ, P φ = P φ</i><small>2</small> <i>≥ 0 for any φ, so the</i>

sum on the right-hand-side of (1.4.8) consists solely of nonnegative terms,

<i>and hence exists as an element of [0,∞]. Let N denote the dimension of V</i>

(which may be finite or infinite), let <i>{ηk}<small>N</small></i>

<i><small>k=1</small></i> be an orthonormal basis of<i>V,</i>

and let <i>{φn} be an orthonormal basis for H. Using Theorem 1.2.14 in the</i>

first line and Parseval’s formula in the penultimate line, one has

<b>Definition 1.4.9. The quantity on the right-hand-side of (1.4.8) is called</b>

<i>the trace of P and is denoted Tr P .</i>

With the basic facts in hand, we now move on to discuss the spectral the-orem for self-adjoint operators on a separable, infinite-dimensional Hilbert space. The situation is more subtle than the finite-dimensional setting. In-deed, we will discuss three results that one can think of as versions of the spectral theorem: the functional calculus, the existence of spectral mea-sures, and the equivalence between self-adjoint operators and multiplication operators.

To motivate what follows and elucidate the subtleties, let us briefly recall the spectral theorem for self-adjoint operators on C<i><small>n</small></i> (i.e., Hermitian

<i>ma-trices). If A∈ C<small>n×n</small><sub>is self-adjoint, then there is a unitary matrix U and a</sub>diagonal matrix D such that U<small>∗</small><sub>AU = D. The columns of U are normalized</sub>eigenvectors of A, and the diagonal entries of D are the corresponding </i>

eigen-values. Now, if <i>H is an infinite-dimensional Hilbert space and A ∈ B(H) isself-adjoint, A may have no eigenvalues whatsoever! The reader is invitedto verify that the discrete Laplacian Δ : </i><small>2</small>(<i>Z) → </i><small>2</small>(Z), given by

(1.4.9) <i>[Δψ](n) = ψ(n + 1) + ψ(n− 1), ψ ∈ </i><small>2</small>(<i>Z), n ∈ Z,</i>

is an example of such an operator (Exercise 1.4.7). That is, Δ<i>∈ B(</i><small>2</small>(Z)), Δ<i><small>∗</small></i> <sub>= Δ, and Δ has no eigenvalues.</sub>

Thus, we need to reformulate the finite-dimensional spectral theorem in a fashion that is amenable to the present situation. To that end, given a

</div><span class="text_page_counter">Trang 15</span><div class="page_container" data-page="15">

<i>self-adjoint matrix A∈ C<small>n×n</small><sub>, write P</sub></i>

<i><small>λ</small></i> for the orthogonal projection onto

<i>Vλ:= Ker(A− λ). Then, the spectral theorem asserts that</i>

<i><small>λ∈σ(A)</small>λP<sub>λ</sub></i>

<i>and that P<sub>λ</sub>Pμ= PμP<sub>λ</sub>= 0 whenever λ = μ. This allows one to see that,for a polynomial function f , f (A) =</i>

<i>f (λ)P<sub>λ</sub>, that is, f (A) depends onlyon the values assumed by f on the spectrum of A.</i>

<i>This is the first point of view that we develop. Suppose that A∈ B(H)is self-adjoint. We wish to apply a function f to A in order to form a newoperator f (A). Recall that if f (z) = c<sub>n</sub>z<sup>n</sup></i>+<i>· · · + c</i><small>1</small><i>z + c</i><sub>0</sub> is a polynomial,

<i>we defined f (A) as follows:</i>

(1.4.11) <i>f (A) = c<sub>n</sub>A<sup>n</sup></i>+<i>· · · + c</i><small>1</small><i>A + c</i><sub>0</sub><i>.</i>

In order to pass from polynomials to more general continuous functions, one would like to understand the behavior of this construction with respect to uniform approximation on compact sets. The following proposition supplies the key input.

(1.4.12) <i>p(A) = sup{|p(z)| : z ∈ σ(A)}.</i>

<b>Proof. This follows from Propositions 1.3.10 and 1.4.7, and the observation</b>

<i>Thus, if A is self-adjoint, it is then possible to define f (A) for functions fthat are continuous on σ(A). Concretely, given f∈ C(σ(A)), let p</i><small>1</small><i>, p</i><sub>2</sub><i>, . . . bea sequence of polynomials that converges uniformly to f on σ(A). Naturally,</i>

one would like to define

The first version of the spectral theorem states that this is well-defined and has the kinds of properties one might desire from such an apparatus.

<i><b>Theorem 1.4.11 (Spectral theorem: continuous functional calculus). Let</b></i>

<i>A∈ B(H) be self-adjoint, and let A ⊆ B(H) denote the algebra generatedby A (that is, the smallest closed subalgebra ofB(H) containing A).</i>

<i>(a) The association from (1.4.13) is well-defined, i.e., the limit existsand does not depend on the choice of approximating polynomials.</i>

</div><span class="text_page_counter">Trang 16</span><div class="page_container" data-page="16">

<i>1.4. Self-Adjoint Operators and the Spectral Theorem</i> 29

<i>(b) The map C(σ(A)) f → f(A) ∈ A is a ∗-homomorphism. Thatis to say, for all continuous f and g and all a∈ C, one has</i>

<i>(f + g)(A) = f (A) + g(A),</i>

<i>where denotes the function that is identically equal to 1 on σ(A).(c) For all f∈ C(σ(A)), f(A) = f<sub>∞</sub>.</i>

<i><small>n=1</small>is a sequence of polynomials with p<sub>n</sub>→ f uniformly onσ(A), then (p<sub>n</sub>(A))<small>∞</small></i>

<i><small>n=1</small></i> is a Cauchy sequence in<i>B(H) by Proposition 1.4.10,</i>

and hence the limit in (1.4.13) exists by completeness of<i>B(H). If (qn</i>)<i><small>∞n=1</small></i>is

<i>another sequence of polynomials that converges uniformly to f on σ(A), thenqn(A) and pn(A) converge to the same limit (again by Proposition 1.4.10).</i>

The proofs of (b) and (c) are left to the reader.  Having constructed the continuous functional calculus, we are now able

<i>to describe the spectral measures of A. Once again, the analogy with the</i>

finite-dimensional case is a useful guide. <i>Let A be a self-adjoint n× ncomplex matrix with spectral decomposition as in (1.4.10). For v∈ C<small>n</small></i> and

<i>f a continuous function, to calculate the corresponding matrix element off (A), we have</i>

<i>f (λ)Pλv</i><small>2</small><i>.</i>

Thus, we can think of the matrix element <i>v, f(A)v as being obtained byintegrating f against the Dirac measure supported on σ(A) with the weightat λ∈ σ(A) given by Pλv</i><small>2</small>. In the infinite-dimensional setting, one can no longer guarantee the existence of a basis of eigenvectors, but one can still speak of the measures defined by the left-hand side of (1.4.19).

<i><b>Theorem 1.4.12 (Spectral theorem: existence of spectral measures). Let</b></i>

<i>A∈ B(H) be self-adjoint. For all φ, ψ ∈ H, there is a complex Borelmeasure μ = μ<sup>A</sup><sub>φ,ψ</sub>such that</i>

</div><span class="text_page_counter">Trang 17</span><div class="page_container" data-page="17">

<i>for all f∈ C(σ(A)). In particular, if ψ = φ, then μ = μ<small>A</small></i>

<i><small>φ</small>:= μ<small>Aφ,φ</small>is ameasure.</i><small>3</small>

<i>φ, f(A)ψ defines a bounded linear functional on C(σ(A)), and hence the</i>

existence of a complex Borel measure satisfying (1.4.20) follows from the Riesz–Markov–Kakutani representation theorem.

<i>Moreover, if φ = ψ and f∈ C(σ(A)) is positive, there is a uniquecontinuous g such that g</i><sup>2</sup><i>= f and g≥ 0. One then has</i>

<i>(f ) =φ, f(A)φ = φ, g(A)</i><small>2</small><i>φ = g(A)φ</i><small>2</small><i>≥ 0,</i>

<i>and hence  is also a positive functional. Thus, in this case μ is also a</i>

<i>The measures constructed in Theorem 1.4.12 are called the spectral mea-sures of A. Whenever A is clear from context, we suppress it from thenotation, writing, e.g., μ<sub>φ,ψ</sub>and μ<sub>φ</sub>for μ<sup>A</sup><sub>φ,ψ</sub>and μ<sup>A</sup><sub>φ</sub></i>.

At this point, let us note a helpful consequence of Theorem 1.4.12 re-garding the norm of the resolvent operator.

<i>dist(z, σ(A))<sup>.</sup></i>

<i>dist(z, σ(A)), and let ψ be a unit vector. By Theorem 1.4.12, we have</i>

proving the desired upper bound on <i>R(A, z). The opposite inequality</i>

holds more generally and follows from Theorem 1.3.6 (see Exercise 1.3.11).  The construction of spectral measures in turn allows one to extend the

<i>map f→ f(A) and the formula (1.4.20) to all bounded Borel measurablefunctions on σ(A). Concretely, given</i>

<i>f∈ B(σ(A)) := {g : σ(A) → C : g is bounded and Borel measurable} ,we define f (A) by specifying its matrix elements,</i>

(1.4.22) <i>φ, f(A)ψ =</i>

<i>f (E) dμ<sup>A</sup><sub>φ,ψ</sub>(E).</i>

<small>3We adopt the convention in this text that “measure” always means “nonnegative regularBorel measure”, using modifiers (e.g., “signed” or “complex”) to distinguish variants.</small>

</div><span class="text_page_counter">Trang 18</span><div class="page_container" data-page="18">

<i>1.4. Self-Adjoint Operators and the Spectral Theorem</i> 31

<i>B(H) be self-adjoint.</i>

<i>(a) For each f∈ B(σ(A)), (1.4.22) defines an element f(A) ∈ B(H).(b) This mapping is a∗-homomorphism from B(σ(A)) to B(H) (cf.</i>

<i>Theorem 1.4.11(b)).</i>

<i>(c) Additionally, if f<sub>n</sub>, n≥ 1 is a sequence in B(σ(A)) that is uni-formly bounded in the supremum norm and fn→ f pointwise onσ(A), then fn(A)→ f(A) strongly (i.e., fn(A)ψ→ f(A)ψ for ev-ery ψ∈ H).</i>

<i>By Theorem 1.4.14(b), let us note that if S⊆ R is a Borel set, then χS</i> =

<i>χ</i><sup>2</sup><i><sub>S</sub>= χS</i> implies that <i>P(S) := χS(A) is an orthogonal projection, knownas the spectral projection of A on S. Here, χ<sub>S</sub>denotes the characteristicfunction of S, given by χ<sub>S</sub>(x) = 1 for x∈ S and 0 otherwise. Using the</i>

Borel functional calculus and the Cauchy–Schwarz inequality, one can check that

<i>|μφ,ψ(S)| = |φ, χS(A)ψ| ≤ μφ(S)<sup>1/2</sup>μ<sub>ψ</sub>(S)<sup>1/2</sup>,</i>

<i>for all Borel S⊆ R, so μφ,ψ</i> is absolutely continuous with respect to both

<i>μφand μψ. The family of these projections comprises a projection-valued-measure (or a resolution of the identity) in thatP(∅) = 0, P(R) = I,P(S ∩ T ) = P(S)P(T ), and, if {Sn: n∈ N} is a countable collectionof pairwise disjoint Borel sets and S =</i> 

Let us discuss one final manifestation of the spectral theorem. To

<i>mo-tivate this, consider again a self-adjoint A∈ C<small>n×n</small></i><sub>. For simplicity, assume</sub> <i>that every eigenvalue of A is simple (this is not necessary, but it makes this</i>

discussion more transparent). Consider the space

<i>H = </i><small>2</small><i>(σ(A)) ={f : σ(A) → C}</i>

<i>and map v∈ C<small>n</small>to f∈ H via f(λ) = eλ, v, where eλ</i> is a normalized

<i>eigenvector corresponding to the eigenvalue λ. Then, the spectral theoremfor finite matrices establishes that the map U taking v to f as above isunitary and [U AU<small>∗</small><sub>g](λ) = λg(λ). Our final instance of the spectral theorem</sub></i>

generalizes this perspective.

</div><span class="text_page_counter">Trang 19</span><div class="page_container" data-page="19">

<b>Theorem 1.4.15 (Spectral theorem: equivalence to multiplication </b>

<i><b>opera-tors). Suppose A</b>∈ B(H) is self-adjoint. Then, there exist at most countablymany φ<sub>n</sub>∈ H such that with μn= μ<sub>φ</sub><sub>n</sub>, there is a unitary map</i>

<i>Here we write g(E) ={gn(E)}nfor an element g∈</i><sup></sup><i><sub>n</sub>L</i><sup>2</sup>(<i>R, μn).</i>

The central notion here is the decomposition of<i>H into cyclic subspaces.</i>

<i>generated by A and φ is the smallest closed subspace ofH containing{A<small>n</small>φ : n = 0, 1, 2, . . .}. An element φ of H is called cyclic for the oper-ator A if the corresponding cyclic subspace is all ofH.</i>

<i>and φ is a cyclic vector for A. We wish to show that (1.4.24) can be definedwith just the measure μ = μ<sub>φ</sub>. To define a map U :H → L</i><small>2</small>(<i>R, μφ</i>) with the

<i>desired property, one needs to invert the map f→ f(A) discussed above.Namely, if f∈ C(σ(A)), set U(f(A)φ) = f; the reader should check thatthis is well-defined (i.e., if f (A)φ = g(A)φ for f, g∈ C(σ(A)), then f = g).This defines U on a dense set. One can then verify that U preserves normsand that (U AU<small>−1</small><sub>f )(E) = Ef (E) for such f ’s and then extend the definition</sub></i>

and the relation to the whole space.

This gives the statement of the spectral theorem if there is a cyclic vector. If not, a Zornification shows that one can decompose <i>H =</i> <sup></sup><i>Hn</i>

where each <i>Hnis a cyclic subspace generated by A and some φ<sub>n</sub>∈ H.Denote μ<sub>n</sub>= μ<sup>A</sup><sub>φ</sub></i>

<i><small>n</small>the spectral measure and U<sub>n</sub></i>:<i>H → L</i><small>2</small>(<i>R, μn</i>) the unitary

<i>equivalence f (A)φ<sub>n</sub>→ f. Then U =</i><sup></sup><i>U<sub>n</sub></i> is the desired unitary map.  The multiplication operators encountered in Theorem 1.4.15 can be thought of as generalizations of direct sums. Let us make this explicit. For this discussion, we only state theorems and leave proofs to the exercises.

<i>a fixed separable Hilbert space. A function A : Ω→ B(H) will be calledweakly measurable if for all φ, ψ∈ H, the function gφ,ψ</i> : Ω<i>→ C,</i>

(1.4.25) <i>g<sub>φ,ψ</sub>: ω→ φ, Aωψ</i>

is measurable; following a convention for operator-valued functions that will

<i>be standard later in the text, we write Aωrather than A(ω) for the value</i>

</div><span class="text_page_counter">Trang 20</span><div class="page_container" data-page="20">

<i>1.4. Self-Adjoint Operators and the Spectral Theorem</i> 33

<i>of A at ω. Similarly, a function ϕ : Ω→ H is said to be measurable</i><small>4</small> if

<i>ψ, ϕ(ω) is measurable for all ψ ∈ H.</i>

<i>The (constant-fiber) direct integral , denoted by L</i><sup>2</sup><i>(Ω, μ;H) or sometimes</i>

<i>consists of measurable maps f : Ω→ H (as always, modulo the subspace of</i>

measurable functions that vanish almost everywhere) for which

<i>L</i><small>2</small><i>(Ω, μ;H) is a Hilbert space. An operator A ∈ B(L</i><small>2</small><i>(Ω, μ;H)) is calleddecomposable if there is a weakly measurable family{Aω} such that</i>

<i>A·∞</i>:= esssup

<i><small>ω∈Ω</small>Aω < ∞</i>

<i>(recall that esssup denotes the μ-essential supremum, defined in (1.2.11))</i>

(1.4.29) <i>[Af ](ω) = Aω(f (ω)),μ-a.e. ω∈ Ω.</i>

In analogy to the notation from (1.4.26), we will sometimes write this as

 <i><sub>⊕</sub></i>

<i>Let us make a few observations about this construction. First, if μ is</i>

the counting measure on Ω, then we see that

the direct sum of copies of <i>H (cf. Definition 1.2.15), and the decomposable</i>

operators are precisely those that can be expressed as direct sums of opera-tors in<i>H (cf. Definition 1.3.12). Thus, the direct integral can be viewed as</i>

a generalization of the direct sum.

<small>4For consistency, we should also call this notion “weak measurability” in order to distinguishit from the ostensibly stronger notion of measurability given by equipping</small><i><small>H with the σ-algebra</small></i>

<small>generated by the metric topology thereupon. However, these two notions are equivalent (Exer-cise 1.12.1).</small>

</div><span class="text_page_counter">Trang 21</span><div class="page_container" data-page="21">

Given a measurable family <i>{Aω}ω<sub>∈Ω</sub></i> with <i>A·∞<∞, one can check(Exercise 1.4.13) that there is a unique decomposable operator A for whichone has (1.4.29) and the operator norm of A satisfies</i>

Let us conclude with a versatile criterion for detecting the spectrum of

<i>a self-adjoint operator. On the one hand, every eigenvalue of an operator Abelongs to the spectrum of A. On the other hand, as we noted earlier, there</i>

are self-adjoint operators (such as Δ, the discrete Laplacian) that have no eigenvalues at all (Exercise 1.4.7). However, for self-adjoint operators, one has the following extremely helpful replacement: one can characterize the

<i>spectra of self-adjoint operators as precisely those z for which the operatorenjoys a sequence of approximate eigenvectors. Let us make this precise.</i>

<i>a sequence (ψ<sub>n</sub></i>)<i><small>∞</small></i>

<i><small>n=1</small>where ψ<sub>n</sub>∈ H for all n, ψn = 1 for all n, and</i>

<i><small>n→∞</small>(A − z)ψn = 0.</i>

<i>z∈ σ(A). If A is self-adjoint, the converse holds as well.</i>

</div><span class="text_page_counter">Trang 22</span><div class="page_container" data-page="22">

<i>1.4. Self-Adjoint Operators and the Spectral Theorem</i> 35

1 =<i>ψn = (A − z)<small>−1</small><sub>(A</sub>− z)ψn ≤ (A − z)<small>−1</small>(A − z)ψn.</i>

Thus, <i>(A − z)ψn ≥ R(A, z)<small>−1</small><sub>and hence (ψn)</sub><small>∞</small></i>

<i><small>n=1</small></i> cannot be a Weyl

<i>sequence for A at z.</i>

<i>Conversely, if E∈ σ(A) and A = A<small>∗</small><sub>, then E</sub>∈ R and E + iδ ∈ ρ(A)for all δ > 0. Choose a sequence δn↓ 0. By Theorem 1.4.13, we can finda sequence (φn)n<small>≥1</small></i> of unit vectors such that <i>(A − E − iδn)<small>−1</small>φn → ∞ as</i>

Let us note the following helpful reformulation.

(1.4.36) <i>dist(z, σ(A)) = inf</i>

<i><small>
v
=1</small>(A − z)v.</i>

<i>right-hand side vanishes by Theorem 1.4.20. If z∈ ρ(A), the claim follows by </i>

not-ing that the left-hand side is equivalent to <i>R(A, z)<small>−1</small></i> <sub>by Theorem 1.4.13</sub>

and the right-hand side is equivalent to the same quantity by a direct

Additionally, we can use the Weyl sequence formalism to bound spectra of strong limits from above.

<i>is a unit vector ψ for which(A − E)ψ < ε. By strong convergence, we get(An− E)ψ < ε</i>

<i>for all large n. For such n, this means that dist(E, σ(A<sub>n</sub>)) < ε by Corol-lary 1.4.21. Consequently, E belongs to the right-hand side of (1.4.37).</i>  We caution the reader that the previous bound is one-sided. In

<i>gen-eral, σ(A) can be much smaller than the right-hand side of (1.4.37) </i>

(Exer-cise 1.4.12).

</div><span class="text_page_counter">Trang 23</span><div class="page_container" data-page="23">

Let us note that there is a characterization of the essential spectrum

<i>via Weyl sequences comprised of orthonormal vectors similar to that of</i>

Theorem 1.4.20.

<i>A at z is an orthonormal sequence of vectors (ψn)<small>∞</small></i>

<i><small>n=1</small>, where ψn∈ H for alln,ψn = 1 for all n, and</i>

<i><small>n→∞</small>(A − z)ψn = 0.</i>

<i>is a singular Weyl sequence for A at E.</i>

<i>multiplic-ity, then one can simply take (ψ<sub>n</sub></i>)<i><small>∞</small></i>

<i><small>n=1</small></i> to be an orthonormal basis of the corresponding eigenspace. Otherwise, with

<i>one has Tr (Pn)≥ 1 for infinitely many n ≥ 1. Thus, the desired orthonormalsequence can be extracted by choosing a normalized vector from Ran(Pn)for each n with Tr (P<sub>n</sub></i>)<i>≥ 1.</i>

<i>Conversely, assume that (ψn)<small>∞</small></i>

<i><small>n=1</small>is a singular Weyl sequence for A at E.By Theorem 1.4.20, E∈ σ(A), so it suffices to show that E /∈ σ</i><small>disc</small><i>(A). As-sume to the contrary that E∈ σ</i><small>disc</small><i>(A), which implies P<sub>ε</sub>:= χ<sub>(E</sub><small>−ε,E+ε)</small><sup>(A)</sup>is a finite-rank projection for some ε > 0. By a direct calculation usingorthonormality, P<sub>ε</sub>ψ<sub>n</sub>→ 0 as n → ∞. On the other hand, writing μψ<small>n</small></i> for

<i>the spectral measure associated to the pair (A, ψn), one has</i>

which is incompatible with <i>ψn = 1 and Pεψn→ 0. Thus, we obtain a</i>

This has the following consequence: the essential spectrum is invariant under finite-rank perturbations.<sup>5</sup>

<i><b>Corollary 1.4.25. If A and R are self-adjoint and R is finite-rank, then</b></i>

(1.4.39) <i>σ</i><sub>ess</sub><i>(A + R) = σ</i><sub>ess</sub><i>(A).</i>

<small>5In fact, the essential spectrum is invariant under compact self-adjoint perturbations, butwe do not state that result since we have not discussed compact operators in the text. Moreover,there are further generalizations beyond the self-adjoint case.</small>

</div><span class="text_page_counter">Trang 24</span><div class="page_container" data-page="24">

<i>1.4. Self-Adjoint Operators and the Spectral Theorem</i> 37

<i>σ</i><sub>ess</sub><i>(A + R) when R is rank-one, say R = λφ, ·φ for some λ and someφ∈ H. Given E ∈ σ</i><small>ess</small><i>(A), Theorem 1.4.24 implies the existence of a singu-lar Weyl sequence (ψ<sub>n</sub></i>)<i><small>∞</small></i>

<i><small>n=1</small>for A at E. By a direct calculation, we have(A + R− E)ψn= (A− E)ψn+ λφ, ψnφ → 0.</i>

<i>The first term goes to zero because the ψncomprise a Weyl sequence for Aat E. The second term goes to zero by orthonormality. Thus, (ψ<sub>n</sub></i>)<i><small>∞</small></i>

<i><small>n=1</small></i> is

<i>also a singular Weyl sequence for A + R at E and hence E∈ σ</i><small>ess</small><i>(A + R),</i>

<i>let φ = (χ)χ with χ a unit vector in Ker()<small>⊥</small><sub>. Show that  = </sub><sub>φ</sub></i><sub>. Conclude</sub> <i>the proof of Proposition 1.4.2 by showing that ψ→ ψ</i> is an anti-linear isometry that maps <i>H onto H<small>∗</small><sub>. Hint : For any ψ, (ψ)φ</sub>− (φ)ψ ∈ Ker().</i>

<b>Exercise 1.4.2. Prove Proposition 1.4.5.</b>

<i>e<small>−itA</small><sub>is unitary for every t</sub>∈ R.</i>

<i>σ(S) = ∂</i>D.

<b>Exercise 1.4.7. Let Δ be the operator defined in (1.4.9).</b>

<i>(a) Show that Δ is a bounded, self-adjoint operator with σ(Δ) =</i>

[<i>−2, 2].</i>

(b) Show that Δ has no eigenvalues.

<i>(c) Let μ = μ</i><sup>Δ</sup><i><sub>δ</sub></i><sub>0</sub> denote the spectral measure associated with the pair <i>Hint : With S as in Exercise 1.4.6, Δ = S + S<small>−1</small></i><sub>.</sub>

</div><span class="text_page_counter">Trang 25</span><div class="page_container" data-page="25">

<b>Exercise 1.4.8. If</b> <i>H is separable and A ∈ B(H) is self-adjoint, show thatA may have at most countably many eigenvalues.</i>

<i>and [Bψ]<sub>n</sub>= ψ<sub>n</sub><small>−1</small><sup>(with the convention ψ</sup></i><small>0</small> = 0).

<i>(a) Show A = B<small>∗</small><sub>, AB = I, and BA</sub> = I.</i>

<i>(b) Show that every z∈ D := {w ∈ C : |w| < 1} is an eigenvalue of A.(c) Show that σ(A) = σ(B) =</i>D.

<i>(d) Show that there is no Weyl sequence for B at 0.</i>

<i>(e) Construct X, Y∈ B(H) such that XY = I and Y X is an </i>

orthog-onal projection onto a subspace of infinite codimension.

<i><b>Exercise 1.4.10. Prove Theorem 1.4.14. Hint : For part (c), use dominated</b></i>

<i><b>Exercise 1.4.11. Following the notation in Exercise 1.3.6, when is D </b></i>

self-adjoint? Unitary? Normal? An orthogonal projection?

<i>K, show that there exist self-adjoint operators Anand A such that An→A strongly, σ(An) = K for all n, and σ(A) ={a}. Hint: Try diagonaloperators on </i><sup>2</sup>(N).

<i>show that there is a unique decomposable operator A for which one has</i>

(1.4.29) and (1.4.32).

<b>Exercise 1.4.14. Prove Theorem 1.4.18.</b>

<i><b>Exercise 1.4.15 (Unbounded operators). Let (D(A), A) be a linear </b></i>

opera-tor in a Hilbert space <i>H for which D(A) is dense in H (in this case, we sayA is densely defined ). Notice that we do not assume that A is bounded.</i>

(a) Define

<i>D(A<small>∗</small></i><sub>) =</sub><i>{φ ∈ H : ψ → φ, Aψ</i>

extends to a bounded linear functional on<i>H}.Prove that, for φ∈ D(A<small>∗</small><sub>), there is a unique element A</sub><small>∗</small><sub>φ</sub>∈ H with</i>

the property that

<i>A<small>∗</small>φ, ψ = φ, Aψ</i>

<i>and that this determines a linear operator (D(A<small>∗</small><sub>), A</sub><small>∗</small></i><sub>) in</sub> <i>H.(b) Show by example that D(A<small>∗</small></i><sub>) =</sub><i>{0} is possible. Hint: Heuristically,</i>

<i>A<small>∗</small></i> <sub>functions like the conjugate transpose of a matrix. Use this as</sub>

a starting point for a precise construction.

</div><span class="text_page_counter">Trang 26</span><div class="page_container" data-page="26">

<i>1.5. Dual Spaces and Locally Convex Topologies</i> 39

<i>(c) Consider the linear operator A defined by D(A) ={φ ∈ </i><small>2</small>(Z) :

<i>supp φ is compact} and [Aφ](n) = nφ(n). Calculate A<small>∗</small></i><sub>.</sub>

<i>(d) One says that A is self-adjoint if (D(A), A) = (D(A<small>∗</small><sub>), A</sub><small>∗</small></i><sub>) </sub> <i>(typi-cally abbreviated simply as A = A<small>∗</small></i><sub>). Find a self-adjoint extension</sub> <i>of the operator A from part (c), that is, a self-adjoint operator(D(B), B) such that D(B)⊃ D(A) and B|D(A)= A.</i>

<b>Notes and Remarks. Proposition 1.4.2 is due to F. Riesz [205] and is</b>

sometimes called the Riesz representation theorem. The spectral theorem

<b>for bounded operators goes back to Hilbert [122], and the extension tounbounded operators is due to von Neumann [262].</b>

The exposition in this section draws inspiration from Reed and Simon

<b>[196], Simon [243], and Teschl [257].</b>

<b>1.5. Dual Spaces and Locally Convex Topologies</b>

We conclude the discussion of functional analysis in general spaces with a brief excursion into dual spaces and alternative topologies that one can place on vector spaces. For ease of exposition and since it suffices for our purposes,

<i>we restrict our attention to real vector spaces and real linear functionals in</i>

the present section.

We say that <i>V is a topological vector space if the topology is Hausdorff and</i>

the natural vector space operations are continuous. That is,

We call<i>V a locally convex topological vector space if in addition there is</i>

a neighborhood basis for the topology at 0 comprised of open convex sets. If <i>V is a normed linear space, the collection N = {B(0, r) : r > 0} is a</i>

neighborhood basis for the norm topology at the origin. In particular, since balls are convex, every normed vector space is locally convex in the norm topology.

Let us briefly recall some terminology and notions from topology. If

<i>T</i><small>1</small> and <i>T</i><small>2</small> <i>are topologies on a set X, we say thatT</i><small>1</small> <i>is weaker thanT</i><small>2</small>

(equivalently, <i>T</i><small>2</small> <i>is stronger thanT</i><small>1</small>) if every set open with respect to

<i>T</i><small>1</small> is open with respect to <i>T</i><small>2</small>. Since the intersection of any collection of

</div><span class="text_page_counter">Trang 27</span><div class="page_container" data-page="27">

topologies is itself a topology, we can in certain situations precisely speak of

<i>the weakest topology enjoying specified properties. For instance, if X is a</i>

set and <i>{fα: α∈ A } is a collection of functions from X into a topologicalspace Y , one can construct on X the weakest topology for which every f<sub>α</sub></i>

is continuous. The locally convex topologies of interest are those that are generated in precisely this fashion by families of linear functionals.

set of linear functions from <i>V to R. We define T (V, L) to be the weakest</i>

topology on <i>V in which every  ∈ L is continuous; equivalently, T (V, L) is</i>

the intersection of all topologies for which every element of <i>L is continuous.</i>

To construct the topology <i>T = T (V, L), we note that a set U ⊆ V is open</i>

with respect to <i>T if and only if for every x ∈ U, there are finitely many</i><small>1</small><i>, . . . , n∈ L and intervals (aj, bj) such that</i>

We say that <i>L is a separating family of linear functionals if for anyu = v in V, there is some  ∈ L such that (u) = (v). Naturally, if L is</i>

a separating family of linear functionals on <i>V, then T (V, L) is a Hausdorff</i>

topology on <i>V.</i>

In the particular case in which <i>V is a Banach space and L = V<small>∗</small></i><sub>, we</sub>

call<i>T (V, V<small>∗</small><sub>) the weak topology on</sub>V. In the case V = U<small>∗</small></i> <sub>for some normed</sub>

space <i>U, each u ∈ U induces a linear functional on V via Eu(v) = v(u).</i>

Identifying <i>U with the collection of evaluation functionals {Eu: u∈ U}, we</i>

call<i>T (U<small>∗</small><sub>,</sub>U) the weak-∗ topology on V.</i>

weak-<i>∗ topology on V<small>∗</small></i> <sub>are locally convex.</sub>

<i>functionals, then the topologyT (V, L) makes V a locally convex topologicalvector space.</i>

<i>n∈ N} be an orthonormal basis of H. Sinceψ</i><small>2</small>=

<i>|φn, ψ|</i><small>2</small><i><∞,</i>

we have <i>φn, ψ → 0. In particular, φn→ 0 weakly. However, </i>

orthonor-mality implies that <i>φn− φm =<sup>√</sup>2 whenever n = m, so (φn)<small>∞</small></i>

<i><small>n=1</small></i> is not convergent in the norm topology.

</div><span class="text_page_counter">Trang 28</span><div class="page_container" data-page="28">

<i>1.5. Dual Spaces and Locally Convex Topologies</i> 41

<i>x<sub>n</sub>→ 0, equipped with the supremum norm. One can identify U<small>∗</small></i> <sub>with</sub> <i>V = </i><small>1</small>(<i>N) and V<small>∗</small><sub>with </sub><small>∞</small></i><sub>(</sub><i>N) (Exercise 1.5.2). One can check that δnconverges to 0 as n→ ∞ in the weak-∗ topology on </i><small>1</small>(N), but not in the

<i>weak topology on </i><sup>1</sup>(<i>N). In fact, one can show that a sequence (yn)<small>∞n=1</small></i> of

<i>elements of </i><sup>1</sup>(N) converges in the weak topology if and only if it converges in the norm topology (Exercise 1.5.3).

<i>C(X,</i>R) be the space of continuous functions equipped with the supremum norm. By the Riesz–Markov–Kakutani representation theorem, <i>V<small>∗</small></i> <sub>may be</sub>

isometrically identified with<i>M(X), the space of signed measures on X with</i>

the total variation norm. In view of these identifications, the weak-<i>∗ </i>

topol-ogy on <i>M(X) is then precisely the weakest topology for which every linear</i>

functional of the form

<i>f dμ</i>

is continuous.

One advantage afforded by the weak-<i>∗ topology is that it makes the</i>

closed unit ball of the dual space a compact space that is metrizable when-ever <i>V is separable. First, we prove compactness.</i>

<i>is compact in the weak-∗ topology.</i>

<i>∈ B to an element x = (xv)v<small>∈V</small>∈ S =</i> <sup>!</sup><i><sub>v</sub><sub>∈V</sub></i>[<i>−v, v], namely bychoosing the element x∈ S whose vth coordinate is (v). Equipping S withthe product topology, S is compact by Tychonoff’s theorem. The mappingf taking ∈ B to x ∈ S is a homeomorphism from B onto its image, by</i>

the definitions of weak-<i>∗ topology and product topology. One can confirmthat the image of B under f is closed. Thus, B is homeomorphic to a closed</i>

subset of a compact space and hence is itself compact. 

Next, we discuss metrizability of the unit ball, which holds whenever the initial Banach space is separable.

<i>ofV<small>∗</small><sub>is metrizable in the weak-</sub>∗ topology.</i>

</div><span class="text_page_counter">Trang 29</span><div class="page_container" data-page="29">

<i><b>Proof. Let S =</b>{vj: j∈ N} be a dense subset of V. The reader can verifythat the function d given on the closed unit ball ofV<small>∗</small></i> <sub>by</sub>

is a metric that generates the weak-<i>∗ topology thereupon.</i> 

<i>is a compact metrizable space in the weak topology.</i>

<b>Proof. This follows by combining Theorems 1.5.8 and 1.5.9 with </b>

Proposi-tion 1.4.2 (which implies that the weak and weak-<i>∗ topologies on H </i>

The next fundamental result we want to prove is the Krein–Milman theorem, which shows that a compact convex set is precisely the closed convex hull of its extreme points. First, we need a definition.

convex. A nonempty subset <i>E ⊆ K is called an extreme subset of K if it isconvex and whenever ax + (1−a)y ∈ E with 0 < a < 1 and x, y ∈ K, one hasx, y∈ E. A point x ∈ K is an extreme point of K if {x} is an extreme subset</i>

of <i>K. The convex hull of a set S ⊆ V is the smallest convex set containingS, that is, the intersection of all convex sets containing S. Similarly, theclosed convex hull ofS is the smallest closed convex set containing S.</i>

<i>locally convex spaceV.</i>

(a) <i>K has at least one extreme point.</i>

(b) <i>K is the closed convex hull of its extreme points.</i>

To prove this, it will be helpful to develop a bit of theory regarding continuous linear functionals on a locally convex space.

If <i>K ⊆ V contains 0 and is nonempty, open, and convex, the gauge of K is</i>

defined by

(1.5.5) <i>p<small>K</small>(v) = inf{c ∈ (0, ∞) : v/c ∈ K}.</i>

One can view this as a generalization of the notion of a norm, for, if<i>V is</i>

a normed space and<i>K is the open unit ball in V, then one has pK(x) =xfor all x∈ V.</i>

</div><span class="text_page_counter">Trang 30</span><div class="page_container" data-page="30">

<i>1.5. Dual Spaces and Locally Convex Topologies</i> 43

<i>convex, and let p = p<small>K</small><sup>denote its gauge. Then p has the following properties.</sup></i>

(a) 0<i>≤ p(v) < ∞ for all v ∈ V.(b) p(x) < 1 for all x∈ K.</i>

<i>(c) p(z)≥ 1 for all z /∈ K.</i>

<i>(d) p(av) = ap(v) for all a≥ 0 and v ∈ V.(e) p(u + v)≤ p(u) + p(v) for all u, v ∈ V.</i>

<i>(f) If  :V → R is a linear functional, then  is continuous if and onlyif ≤ p<sub>K</sub>for some nonempty, open, convexK containing 0.</i>

<i>any open, convexK ⊆ V and any y ∈ V \ K, there is a continuous linearfunctional  onV such that</i>

<i>Since y /∈ K, we have p(y) ≥ 1. In particular, if we define a linearfunctional on the span of y by (ay) = a, we have</i>

<i>By Lemma 1.5.14, p is positive homogeneous and subadditive, so we may</i>

apply the Hahn–Banach theorem (cf. Exercise 1.3.13) to construct a linear

<i>functional  for which (y) = 1 and (v)≤ p(v) for all v ∈ V. In particular,(x)≤ p(x) < 1 = (y) for every x ∈ K. By Lemma 1.5.14,  is continuous,</i>

so the first statement is proved.

<i>For the second statement, suppose x = y. By local convexity, there isan open convex neighborhood of x that does not meet y. Apply the firstpart of the lemma to this set and y yields a continuous linear functional </i>

We can push just a bit further and also get a stronger separation

<i>state-ment for closed convex sets.</i>

</div><span class="text_page_counter">Trang 31</span><div class="page_container" data-page="31">

<i><b>Corollary 1.5.16. Suppose</b>V is a locally convex topological vector space.For any closed, convexK ⊆ V and any y ∈ V \ K, there is a continuouslinear functional  onV and a scalar α such that</i>

(1.5.8) <i>(x) < α < (y)for all x∈ K.</i>

topology of <i>V is locally convex, there is an open convex set G containing</i>

0 that is disjoint from <i>K. In particular, the set K</i><small>1</small> = <i>K − G is open and</i>

convex (cf. Exercise 1.5.4) and does not contain 0, so Lemma 1.5.15 implies

<i>that there is a linear functional  with (x) < (y) = (0) = 0 for all x∈ K</i><small>1</small>.

<i>Since  does not vanish identically andG is a neighborhood of 0, there existsz∈ G with (z) = α < 0. Since (x) < 0 for all x ∈ K</i><small>1</small><i>, we get for all w∈ K</i>

(1.5.9) <i>(w) = (z) + (w− z) < α + 0 < 0 = (y),</i>

extreme subsets of <i>K. By compactness, the intersection of any totally </i>

or-dered subcollection of<i>E is nonempty. Thus, by a Zornification, there exists</i>

a minimal element<i>S of E . We claim that S consists of a single point. Sup-pose on the contrary that x</i><sub>1</sub> <i> = x</i><small>2</small> belong to <i>S. By Lemma 1.5.15, thereexists a continuous linear functional  defined onV such that (x</i><small>1</small>)<i> = (x</i><small>2</small>).

<i>By compactness,  attains a maximal value M onS. One can check thatthe set of x∈ S for which (x) = M is an extreme subset of K that must be</i>

strictly smaller than <i>S, since  is nonconstant on S; a contradiction. Thus,S consists of a single point and therefore K has at least one extreme point.</i>

(b) Let <i>E ⊆ K denote the collection of all extreme points of K, and</i>

let <i>K denote the closed convex hull of E. Naturally, K ⊆ K, so it remainsto show the opposite inclusion. To that end, assume that u∈ V \ K. ByCorollary 1.5.16, there is a continuous linear functional  such that (v) <(u) for all v∈ K. By compactness,  attains a maximum value M on K.The set of x∈ K for which (x) = M is a closed extreme subset of K andcontains an element y ofE by a repetition of the Zornification in part (a).In particular, for all z∈ K, one has (z) ≤ M = (y) < (u), showing u /∈ K,</i>

<b>Exercise 1.5.1. Prove Theorem 1.5.4.</b>

</div><span class="text_page_counter">Trang 32</span><div class="page_container" data-page="32">

<i>1.5. Dual Spaces and Locally Convex Topologies</i> 45

<i><small>n=1</small>: xn∈ R, xn→ 0 asn→ ∞} consist of all real-valued sequences converging to zero, equippedwith the <small>∞</small></i> <sub>metric.</sub>

<i>(a) Show that every y∈ </i><small>1</small>(N) defines a bounded linear functional on

<i>(b) Show that this gives an isometric bijection of c</i><sub>0</sub>(N)<i><small>∗</small><sub>with </sub></i><small>1</small>(N).

<i>(c) In a similar fashion, show that <small>∞</small></i><sub>(</sub>N) may be isometrically

<i>identi-fied with </i><sup>1</sup>(N)<i><small>∗</small></i><sub>.</sub>

<i>(d) Show that δn→ 0 in the weak-∗ topology on </i><small>1</small>(N) but not in the weak topology.

<i><small>n=1</small></i> denote a sequence of elements of <i>V. Prove that yn→ y in the weak topology on V if and only if yn→ y</i>

in the norm topology on <i>V. Hints: For the hard direction, assume to thecontrary that yn→ 0 weakly in </i><small>1</small> and <i>yn </i><small>1</small> <i> → 0. Pass to a subsequencez<sub>k</sub>= yn<sub>k</sub>such that “most” of the norm of z<sub>k</sub>is concentrated on a set S<sub>k</sub></i> in

<i>such a way that the S<sub>k</sub></i>’s are disjoint.

(a) If <i>K</i><small>1</small><i>,K</i><small>2</small> <i>⊆ V are convex, show that K</i><small>1</small><i>− K</i><small>2</small>=<i>{x</i><small>1</small><i>− x</i><small>2</small> <i>: x<sub>j</sub>∈ Kj}</i>

is convex. Moreover, if <i>V is a topological vector space, show thatK</i><small>1</small><i>− K</i><small>2</small> is open whenever<i>K</i><small>1</small> is open.

<i>(b) If T :V → W is linear and K ⊆ V is convex, show that T (K) is</i>

<i>(c) If T :V → W is linear and K ⊆ W is convex, show that T<small>−1</small></i><sub>(</sub><i>K) is</i>

<b>Exercise 1.5.5. Prove Lemma 1.5.14.</b>

set of linear functions from<i>V to R. If U ⊆ V is open with respect to T (V, L)</i>

and 0<i>∈ U, show that there are linearly independent </i><small>1</small><i>, . . . , n∈ L and δ > 0</i>

set of linear functions from<i>V to R. Show that a linear functional  : V → R</i>

is continuous with respect to the topology <i>T (V, L) if and only if  is a</i>

finite linear combination of elements of <i>L. Hint: Start by showing that if</i>

</div><span class="text_page_counter">Trang 33</span><div class="page_container" data-page="33">

<i>{</i><small>1</small><i>, . . . , n} is a linearly independent subset of L, then the range of the mapv→ (</i><small>1</small><i>(v), . . . , <sub>n</sub>(v)) is all of</i> R<i><small>n</small></i>.

<b>Notes and Remarks. The exposition in this section is inspired by Lax</b>

<b>[166] and Simon [244].</b>

Theorem 1.5.8 is usually named the Banach–Alaoglu theorem after

<b>con-tributions of those authors [5,21]. The result was discovered in various formsindependently by many authors. Hilbert [122], Riesz [203], and Helly [120]had special cases before the general result, while Bourbaki [44], Shmulyan[228], and Kakutani [137] proved general results independently around the</b>

same time as Alaoglu. Discussing the number of authors with theorems related to this result, Pietsch remarks:

To be historically complete, one should speak of the Ascoli–Hilbert–Fr´echet–Riesz–Helly–Banach–Tychonoff– Alaoglu–Cartan–Bourbaki–Shmulyan–Kakutani theorem.

<b>Theorem 1.5.12 is due to Krein and Milman [157].</b>

<b>1.6. Spectral Decompositions and Quantum Dynamics</b>

Quantum mechanics suggests the study of the Schrăodinger equation, which is a time-dependent problem. A classical and very important approach to this problem is based on a detailed spectral analysis of the Schrăodinger operator appearing in the Schrăodinger equation, which is in fact a time-independent problem. This has led to the development of the study of suitable spectral decompositions, which may be investigated for any self-adjoint operator. These decompositions can then be connected to corresponding behaviors in the time-dependent setting. In this section we develop the traditional aspects of this approach and present the decomposition of spectral measures and spectra into their absolutely continuous, singular continuous, and pure point parts, and we furthermore discuss how each piece of the decomposition gives information about the behavior of the solution of the time-dependent Schrăodinger equation via the so-called RAGE theorem. After this section, the next two sections address refinements in the pure point and singular continuous subspaces.

<b>1.6.1. Decompositions of Measures and Operators. Recall that any</b>

<i>measure μ on</i> R has a unique Lebesgue decomposition (1.6.1) <i>μ = μ</i><sub>pp</sub><i>+ μ</i><sub>sc</sub><i>+ μ</i><sub>ac</sub><i>,</i>

<i>where μ</i><small>pp</small><i>is a pure point measure, μ</i><small>sc</small>is a singular continuous measure, and

<i>μ</i><small>ac</small> <i>is an absolutely continuous measure. This means that μ</i><small>pp</small>(<i>R \ C) = 0</i>

</div><span class="text_page_counter">Trang 34</span><div class="page_container" data-page="34">

<i>1.6. Spectral Decompositions and Quantum Dynamics</i> 47

<i>for some countable set C, μ</i><small>sc</small>(<i>{E}) = 0 for every E ∈ R, μ</i><small>sc</small>(<i>R \ S) = 0for some set S of zero Lebesgue measure, and μ</i><sub>ac</sub><i>(N ) = 0 for every set Nof zero Lebesgue measure. We also write μ</i><sub>c</sub> <i>= μ</i><sub>sc</sub><i>+ μ</i><sub>ac</sub> <i>for the continuouspart and μ</i><sub>s</sub> <i>= μ</i><sub>pp</sub><i>+ μ</i><sub>sc</sub> <i>for the singular part of μ.</i>

<i>In general, we write μ ν to mean μ is absolutely continuous withrespect to ν, that is, μ(S) = 0 whenever ν(S) = 0. If μ and ν are mutuallyabsolutely continuous, that is, μ ν and ν μ, we say that μ and ν areequivalent. We write μ⊥ ν to mean μ and ν are mutually singular, that is,there are disjoint sets A and B such that μ(R \ A) = ν(R \ B) = 0.</i>

is self-adjoint. Define the following spaces:

Whenever we can make statements about multiple pieces of the spectral decomposition simultaneously, we may employ<i>• as a stand-in for ac, sc, pp,</i>

c, and s. Concretely, we could rephrase the definitions above by declaring (1.6.7) <i>H<sub>•</sub></i> =<i>{φ : μφ= μ<sub>φ,</sub><small>•</small>}, • ∈ {ac, sc, pp, c, s}.</i>

<i>(a) The setsH<sub>•</sub>,• ∈ {ac, sc, pp, c, s}, are A-invariant closed subspaces</i>

<i>{ac, sc, pp, c, s} is closed and A-invariant, so, for each •, we can define alinear operator A<small>•</small><sup>:= A</sup>|H<sub>•</sub>∈ B(H•) by restricting A toH•</i>. Define

(1.6.8) <i>σ<small>•</small>(A) = σ(A<small>•</small>),• ∈ {ac, sc, pp, c, s}.</i>

<i>Respectively, the sets σ</i><sub>ac</sub><i>(A), σ</i><sub>sc</sub><i>(A), σ</i><sub>pp</sub><i>(A), σ</i><sub>c</sub><i>(A), and σ</i><sub>s</sub><i>(A) are called</i>

the absolutely continuous, singular continuous, pure point, continuous, and

<i>singular spectra of A. Finally, letP•</i>=<i>P•(A) be the orthogonal projection</i>

onto <i>H•</i> for<i>• ∈ {ac, sc, pp, c, s}.</i>

</div><span class="text_page_counter">Trang 35</span><div class="page_container" data-page="35">

<i>We say that A has purely absolutely continuous spectrum ifH</i><small>ac</small> = <i>H,purely singular continuous spectrum ifH</i><small>sc</small> = <i>H, pure point spectrum</i><small>6</small> if

<i>H</i><small>pp</small> =<i>H, purely continuous spectrum if H</i><small>c</small>=<i>H, and purely singular spec-trum ifH</i><small>s</small> =<i>H.</i>

<i>or-thogonal, σ</i><sub>ac</sub><i>(A), σ</i><sub>sc</sub><i>(A), and σ</i><sub>pp</sub><i>(A) need not be pairwise disjoint </i>

(Exer-cise 1.6.3).

<i>A nonzero vector φ∈ H is called an eigenvector of an operator A ifthere is a so-called eigenvalue E∈ C with the property Aφ = Eφ. We writeσ</i><sub>p</sub><i>(A) for the set of eigenvalues of A. Eigenvalues are intimately connected</i>

to point spectrum:

<i>(a) Every eigenvector of A belongs toH</i><small>pp</small><i>.(b) Every eigenvalue of A belongs to σ</i><sub>pp</sub><i>(A).</i>

<i>(c) Aψ = Eψ if and only if ψ∈ Ran χ<sub>{E}</sub>(A).</i>

(d) <i>H</i><small>pp</small><i>is the smallest closed linear subspace ofH containing all eigen-vectors of A.</i>

<i>(e) The pure point spectrum of A is given by</i>

<i>E. By induction one has f (A)ψ = f (E)ψ for any polynomial f and henceany continuous f by approximation. Thus, for any continuous f ,</i>

<i>f dμ<sub>ψ</sub></i> =<i>ψ, f(A)ψ = ψ, f(E)ψ = f(E)ψ</i><small>2</small><i>,</i>

<i>showing that μ<sub>ψ</sub></i> is <i>ψ</i><small>2</small> <i>times a Dirac δ measure at E. Thus, μ<sub>ψ</sub></i> is a point

<i>measure, implying ψ∈ H</i><small>pp</small><i>, and in particular that (A</i><sub>pp</sub><i>− E)ψ = 0, whichproves (a) and (b). Moreover, since μψ</i> =<i>ψ</i><small>2</small><i>δE</i>,

(1.6.11) <i>χ<sub>{E}</sub>(A)ψ</i><small>2</small>=<i>ψ, χ<sub>{E}</sub>(A)ψ =</i>

<i>proving that ψ∈ Ran χ<sub>{E}</sub>(A) in this case and one direction of (c). For theother direction, assume ψ∈ Ran χ<sub>{E}</sub>(A) and note that for any φ, one has</i>

(1.6.12) <i>φ, ψ = φ, χ<sub>{E}</sub>(A)ψ =</i>

<i>χ<sub>{E}</sub>dμφ,ψ= μφ,ψ</i>({E}),

<small>6Strict parallelism would insist that we refer to this as “purely pure point spectrum”, butthe usage here is standard.</small>

</div><span class="text_page_counter">Trang 36</span><div class="page_container" data-page="36">

<i>1.6. Spectral Decompositions and Quantum Dynamics</i> 49

which yields in turn

<i>φ, Aψ = φ, Aχ<sub>{E}</sub>(A)ψ =</i>

<i>xχ<sub>{E}</sub>(x) dμ<sub>φ,ψ</sub>(x)= E· μφ,ψ</i>(<i>{E})</i>

=<i>φ, Eψ.</i>

Since <i>φ, Aψ = φ, Eψ for all φ, Aψ = Eψ, concluding the proof of (c).</i>

By (a), <i>H</i><small>pp</small> contains every eigenvector, hence it contains their closed

<i>linear span. Conversely, given ψ∈ H</i><small>pp</small><i>, μ<sub>ψ</sub></i> is a point measure, so we can

<i>which, in view of (c), shows that ψ belongs to the closed linear span of the</i>

eigenvectors, concluding the proof of (d). Then (e) follows from (b) and (d)

<i>since by the spectral theorem, σ</i><small>pp</small><i>(A) is the smallest closed set that supports</i>

<i>If the operator A has a cyclic vector φ, then the spectral theorem assertsthat A is unitarily equivalent to multiplication by the independent variableon L</i><small>2</small>(<i>R, μφ) via the canonical unitary map U : g(A)φ→ g. By choosingdisjoint supports of μφ,pp, μφ,sc, and μφ,ac, we obtain the orthogonal sum</i>

(1.6.14) <i>L</i><sup>2</sup><i>(μφ) = L</i><sup>2</sup><i>(μφ,pp)⊕ L</i><small>2</small><i>(μφ,sc)⊕ L</i><small>2</small><i>(μφ,ac).</i>

Moreover, this decomposition corresponds to the orthogonal splitting <i>H =H</i><small>ac</small><i>⊕ H</i><small>sc</small><i>⊕ H</i><small>pp</small> <i>in the sense that the unitary equivalence U mapsH<sub>•</sub></i> onto

<i>L</i><small>2</small>(<i>R, μφ,<sub>•</sub></i>) for <i>• ∈ {ac, sc, pp}. In particular, we can recover the point,singular continuous, and absolutely continuous spectra of A from supportsof the decomposition of μ<sub>φ</sub>. In general, a support of a measure μ is anyset whose complement has zero μ-measure. We also define the topologicalsupport of μ, supp μ, to be the intersection of all closed supports of μ, orequivalently, the complement of the union of all open sets of zero μ-measure.</i>

(1.6.15) <i>σ(A) = supp(μ<sub>φ</sub>),σ<small>•</small><sup>(A) = supp(μ</sup><small>φ,•</small><sup>),</sup>• ∈ {ac, sc, pp}.</i>

<b>Proof. We write the proof of the first statement and leave the modifications</b>

<i>for the other three to the reader. By the spectral theorem, A is unitarilyequivalent to the operator M : L</i><sup>2</sup><i>(μ<sub>φ</sub></i>)<i>→ L</i><small>2</small><i>(μ<sub>φ</sub>) given by [M g](x) = xg(x)</i>

</div><span class="text_page_counter">Trang 37</span><div class="page_container" data-page="37">

<i>and hence it suffices to show σ(M ) = supp(μ<sub>φ</sub>). If E /∈ supp(μφ</i>), one can

<i>check that M− E is invertible with inverse</i>

so<i>gn<small>−1</small><sub>gn</sub><sub>is a Weyl sequence for M at E, showing E</sub>∈ σ(M).</i>  Let us note that Theorem 1.6.6 suggests how one can go about construct-ing operators with prescribed spectral properties. For instance, to find an operator with singular continuous spectrum equal to a given compact set

<i>K⊆ R, it suffices to construct a finite Borel measure μ on R such thatsupp(μ</i><small>sc</small><i>) = K and to consider the operator [Af ](x) = xf (x) in L</i><sup>2</sup><i>(μ).</i>

<b>1.6.2. The RAGE Theorem. The Lebesgue decomposition of a spectral</b>

<i>measure μφ</i> is also of importance in the study of the long-time asymptotics of solutions to the Schrăodinger equation

(1.6.16) <i>it(t) = A(t), (0) = φ.</i>

Equation (1.6.16) has a unique solution given by

which implies that the evolution is unitary (compare Exercise 1.4.3). The following theorem establishes precise relations between spectral measures

<i>and solutions of (1.6.16) for operators in </i><small>2</small>(Z).

<i>φ∈ </i><small>2</small>(<i>Z), and φ(t) is a solution of the Schrăodinger equation (1.6.16).(a) We have <sub></sub>= <sub>,pp</sub>if and only if for every ε > 0, there is N∈ Z</i><small>+</small>

</div><span class="text_page_counter">Trang 38</span><div class="page_container" data-page="38">

<i>1.6. Spectral Decompositions and Quantum Dynamics</i> 51

The RAGE theorem has its roots in the following celebrated theorem of

<i>Wiener regarding Fourier transforms of measures. The Fourier transformof a finite Borel measure μ on</i> R is defined by

so (1.6.22) follows by the dominated convergence theorem.  We now prove a technical result that will be useful in the proof of the RAGE theorem and also later on in the text. In fact, we will prove the “only

<i>if” direction of (b) in the course of proving this lemma. For N∈ Z</i><small>+</small>, let

<i>PN</i> denote projection onto coordinates in [<i>−N, N]. That is, PNδn= δn</i> for

<i>|n| ≤ N and PNδn</i>= 0 otherwise.

<i>denoteP<sub>•</sub></i> =<i>P<sub>•</sub>(A). Then, we have</i>

<i>proof of (1.6.24). Let φ, ψ be given, and abbreviate φ<small>•</small></i> =<i>P•φ and ψ<small>•</small></i>=<i>P•ψ</i>

for<i>• ∈ {c, pp}.</i>

</div><span class="text_page_counter">Trang 39</span><div class="page_container" data-page="39">

<i>If φ = φ</i><small>pp</small> <i>and ψ = ψ</i><small>pp</small><i>, then, by Proposition 1.6.5, we may write φ andψ as convergent sums of eigenvectors of A. Having done so, we deduce</i>

from a calculation that we leave to the reader (Exercise 1.6.6).

<i>Now, suppose φ = φ</i><sub>c</sub><i>, that is, φ∈ </i><small>2</small>(Z)<small>c</small>. Denote by Π<i><sub>k</sub></i> projection onto

<i>the coordinate k∈ Z so that PN</i> =  <i>Since μδ<sub>k</sub><small>,φ</small>is absolutely continuous with respect to μφ= μφ,c, Wiener’stheorem implies that the final line tends to zero as T→ ∞. The sameargument holds true if ψ∈ </i><small>2</small>(Z)<small>c</small> <i>(and φ is arbitrary). Consequently, for</i>

<i>whenever at least one of φ or ψ is in </i><sup>2</sup>(Z)<small>c</small>. (Notice that this already proves

<i>the “only if” direction of Theorem 1.6.7(b) by taking ψ = φ.)At last, given general φ, ψ, we have</i>

by (1.6.27) and (1.6.26), which completes the proof of (1.6.24). 

<b>Proof of Theorem 1.6.7. (a) For the “only if” direction, suppose we are</b>

<i>given φ∈ </i><small>2</small>(Z)<small>pp</small><i>, and δ > 0. We can choose a finite sum of eigenfunctionsη = η</i><sub>1</sub>+<i>· · · + ηk</i> so that

<i>Aηj= λjηj, andφ − η ≤ δ.</i>

<i>(I − PN)φ(t) = (I − PN)e<small>−itA</small><sub>φ</sub></i>

<i>≤ (I − PN)e<small>−itA</small><sub>η</sub> + (I − PN)e<small>−itA</small><sub>(φ</sub>− η)</i>

</div><span class="text_page_counter">Trang 40</span><div class="page_container" data-page="40">

<i>1.6. Spectral Decompositions and Quantum Dynamics</i> 53

<i>For all N sufficiently large, the first term of the bottom line is no largerthan δ for all t. Consequently,</i>

<i>For the converse, if for every ε > 0 there exists an N so that (1.6.18)holds, then setting ψ = φ in (1.6.25) yieldsP</i><small>c</small><i>φ = 0, whence μ<sub>φ</sub>= μ<sub>φ,pp</sub></i>.

<i>(b) Taking φ = ψ in (1.6.27) proves the “only if” implication, so itremains to show the “if” direction. If (1.6.19) holds for all N , then (1.6.24)</i>

implies <i>P</i><small>pp</small><i>φ = 0, which implies μ<sub>φ</sub>= μ<sub>φ,c</sub></i>.

(c) This follows from the same argument as in the forward direction of (b), but with the Riemann–Lebesgue lemma replacing Wiener’s theorem. 

<i>In the mathematical formulation of quantum mechanics, A is given by</i>

a so-called Schră<i>odinger operator and for a solution (Ã) of the associated</i>

Schrăodinger equation, <i>|δn, φ(t)|</i><small>2</small> corresponds to the probability of finding

<i>the system at site n at time t. Thus, in this situation, we can interpret the</i>

results of the RAGE theorem as follows. The spectral measure of the initial state is pure point if and only if with arbitrarily high probability, the system can be found in some fixed compact region in space at any given time; the spectral measure of the initial state is purely continuous if and only if the system leaves any fixed compact region in space in a time-averaged sense as time goes to infinity; and if the spectral measure of the initial state is purely absolutely continuous, then the system leaves any fixed compact region in space as time goes to infinity.

<b>1.6.3. Dynamical Averages via Resolvent Averages. We conclude</b>

this section with a useful lemma that connects the time-averaged quantum evolution to an integral involving the resolvent function at energies with

<i>constant nonzero imaginary part. For the proof, let us define the Fourier</i>

</div>

×