Spectral Mapping Theorem For Polynomials
Spectral Mapping Theorem For Polynomials
Spectrum
Theorem 2.1. The resolvent set ρ(T ) is nonempty and open, and the spec-
trum σ(T ) is compact .
Proof. Take any T ∈ B[X ]. By the Neumann expansion (Theorem 1.3), if
T < |λ|, then λ ∈ ρ(T ). Equivalently, since σ(T ) = C \ρ(T ),
Remark. Since Bδ (λ) ⊆ ρ(T ), it follows that the distance of any λ in ρ(T ) to
the spectrum σ(T ) is greater than δ; that is (compare with Proposition 2.E),
Since RT (λ) − RT (ν) = RT (λ)[RT (ν)−1 − RT (λ)−1 ]RT (ν), it follows that
f (ν) with the following property. For every ε > 0 there is a δ > 0 such that
f (λ)−f (ν)
− f (ν) < ε for all λ in Λ for which 0 < |λ − ν| < δ. If there exists
λ−ν
such an f (ν) ∈ C , then it is called the derivative of f at ν. If f (ν) exists for
every ν in Λ, then f : Λ → C is analytic (or holomorphic) on Λ. A function
f : C → C is entire if it is analytic on the whole complex plane C . To prove the
next result we need the Liouville Theorem, which says that every bounded
entire function is constant.
Remark. σ(T ) is compact and nonempty, and so is its boundary ∂σ(T ). Hence,
∂σ(T ) = ∂ρ(T ) = ∅.
σ(T ) = σP (T ) ∪ σC (T ) ∪ σR (T ).
2.2 A Classical Partition of the Spectrum 31
The diagram below, borrowed from [62], summarizes such a partition of the
spectrum. The residual spectrum is split into two disjoint parts, σR(T ) =
σR1(T ) ∪ σR2(T ), and the point spectrum into four disjoint parts, σP (T ) =
4
i=1 σPi(T ). We adopt the following abbreviated notation: Tλ = λI − T ,
Nλ = N (Tλ ), and Rλ = R(Tλ ). Recall that if N (Tλ ) = {0}, then its linear
inverse Tλ−1 on Rλ is continuous if and only if Rλ is closed (Theorem 1.2).
R−
λ
=X R−
λ
= X
R−
λ
= Rλ R−
λ
= Rλ R−
λ
= Rλ R−
λ
= Rλ
Theorem 2.2 says that σ(T ) = ∅, but any of the above disjoint parts of
the spectrum may be empty (Section 2.7). However, if σP (T ) = ∅, then a set
of eigenvectors associated with distinct eigenvalues is linearly independent.
There are some overlapping parts of the spectrum which are commonly
used. For instance, the compression spectrum σCP (T ) and the approximate
point spectrum (or approximation spectrum) σAP (T ), which are defined by
σCP (T ) = λ ∈ C : R(λI − T ) is not dense in X
= σP3(T ) ∪ σP4(T ) ∪ σR (T ),
σAP (T ) = λ ∈ C : λI − T is not bounded below
= σP (T ) ∪ σC (T ) ∪ σR2(T ) = σ(T )\σR1(T ).
The points of σAP (T ) are referred to as the approximate eigenvalues of T .
Proof. Clearly (a) implies (b). If (b) holds, then there is no constant α > 0
such that α = αxn ≤ (λI − T )xn for all n. Thus λI − T is not bounded
below, and so (b) implies (c). Conversely, if λI − T is not bounded below,
then there is no constant α > 0 such that αx ≤ (λI − T )x for all x ∈ X
or, equivalently, for every ε > 0 there exists a nonzero yε in X such that
(λI − T )yε < εyε . Set xε = yε −1 yε , and hence (c) implies (a).
If supn (λn I − T )−1 < ∞, then (λn I − T )−1 → (λI − T )−1 in B[X ], be-
cause (λn I − T ) → (λI − T ) in B[X ], and so (λI − T )−1 ∈ B[X ]. That is,
(λI − T ) ∈ G[X ], which is again a contradiction (since λ ∈ σ(T )). This com-
pletes the proof of the claimed result.
Since (λn I − T )−1 = supy=1 (λn I − T )−1 y, take a unit vector yn in
X for which (λn I − T )−1 − n1 ≤ (λn I − T )−1 yn ≤ (λn I − T )−1 for
each n. The above claim ensures that supn (λn I − T )−1 yn = ∞, and hence
inf n (λn I − T )−1 yn −1 = 0, so that there exist subsequences of {λn } and
{yn }, say {λk } and {yk }, for which
Set xk = (λk I − T )−1 yk −1 (λk I − T )−1 yk and get a sequence {xk } of unit
vectors in X such that (λk I − T )xk = (λk I − T )−1 yk −1 . Hence
(λI − T )xk = (λk I − T )xk − (λk − λ)xk ≤ (λk I − T )−1yk −1 + |λk − λ|.
∂σ(T ) ⊆ σAP (T ).
This inclusion clearly implies that σAP (T ) = ∅ (for σ(T ) is closed and non-
empty). Finally, take an arbitrary λ ∈ C \σAP (T ) so that λI − T is bounded
below. Therefore, there exists an α > 0 for which
For the remainder of this section we assume that T lies in B[H], where
H is a nonzero complex Hilbert space. This will bring forth some important
34 2. Spectrum
σR (T ) = σP (T ∗ )∗ \σP (T ).
Proof. Since S ∈ G[H] if and only if S ∗ ∈ G[H], we get ρ(T ) = ρ(T ∗ )∗ . Hence
σ(T )∗ = (C \ρ(T ))∗ = C \ρ(T ∗ ) = σ(T ∗ ). Recall that R(S)− = R(S) if and
only if R(S ∗ )− = R(S ∗ ), and N (S) = {0} if and only if R(S ∗ )⊥ = {0} (cf.
Lemmas 1.4 and 1.5), which means that R(S ∗ )− = H. Thus σP1(T ) = σR1(T ∗ )∗,
σP2(T ) = σR2(T ∗ )∗, σP3(T ) = σP3(T ∗ )∗, and σP4(T ) = σP4(T ∗ )∗. Applying the
same argument, σC (T ) = σC (T ∗ )∗ and σCP (T ) = σP (T ∗ )∗ . Hence,
Therefore, σAP (T ∗ )∗ ∩ σAP (T ) = σ(T )\(σP1(T ) ∪ σR1(T )). But σ(T ) is closed
and σR1(T ) is open (and so is σP1(T ) = σR1(T ∗ )∗ ) in C . This implies that
σP1(T ) ∪ σR1(T ) ⊆ σ(T )◦ , where σ(T )◦ denotes the interior of σ(T ), and
∂σ(T ) ⊆ σ(T )\(σP1(T ) ∪ σR1(T )).
with βn = −1n+1 αn where {λi }ni=1 are the roots of ν − p(λ), so that
n n
νI − p(T ) = νI − αi T i = βn (λi I − T ).
i=0 i=1
If λi ∈ ρ(T ) for every i = 1, . . . , n, then βn ni=1 (λi I − T ) ∈ G[X ] so that
ν ∈ ρ(p(T )). Thus if ν ∈ σ(p(T )), then there exists λj ∈ σ(T ) for some j =
1, . . . , n. However, λj is a root of ν − p(λ), that is,
n
ν − p(λj ) = βn (λi − λj ) = 0,
i=1
since (λj I − T ) commutes with (λi I − T ) for every i. If ν ∈ ρ(p(T )), then
(νI − p(T )) ∈ G[X ] so that
n
(λj I − T ) βn (λi I − T ) νI − p(T ) −1
j=i=1
= νI − p(T ) νI − p(T ) −1 = I = νI − p(T ) −1 νI − p(T )
n
= (νI − p(T ) −1 βn (λi I − T ) (λj I − T ).
j=i=1
Then (λj I − T ) has a right and a left inverse (i.e., it is invertible), and so
(λj I − T ) ∈ G[X ] by Theorem 1.1. Hence, λ = λj ∈ ρ(T ), which contradicts
the fact that λ ∈ σ(T ). Conclusion: if ν ∈ p(σ(T )), then ν ∈/ ρ(p(T )), that is,
ν ∈ σ(p(T )). Thus
p(σ(T )) ⊆ σ(p(T )).
In particular,
σ(T n ) = σ(T )n for every n ≥ 0,
that is, ν ∈ σ(T )n = {λn ∈ C : λ ∈ σ(T )} if and only if ν ∈ σ(T n ), and
that is, ν ∈ ασ(T ) = {αλ ∈ C : λ ∈ σ(T )} if and only if ν ∈ σ(αT ). Also notice
(even though this is not a particular case of the Spectral Mapping Theorem
for polynomials) that if T ∈ G[X ], then
σ(T ∗ ) = σ(T )∗
Take a surjective homomorphism Φ: A(T ) → C (i.e., take any Φ ∈ A(T )). Con-
sider the Cartesian decomposition T = A + iB, where A, B ∈ B[H] are self-
adjoint, and so T ∗ = A − iB (Proposition 1.O). Thus Φ(T ) = Φ(A) + iΦ(B)
and Φ(T ∗ ) = Φ(A) − iΦ(B). Since A = 12 (T + T ∗ ) and B = − 2i (T − T ∗ ) lie in
P(T, T ∗), we get A(A) = A(B) = A(T ). Moreover, since they are self-adjoint,
)} = σ(A) ⊂ R and {Φ(B) ∈ C : Φ ∈ A(T
{Φ(A) ∈ C : Φ ∈ A(T )} = σ(B) ⊂ R
(Propositions 2.A and 2.Q(b)), and so Φ(A) ∈ R and Φ(B) ∈ R . Hence
Φ(T ∗ ) = Φ(T ).
Therefore, since σ(T ∗ ) = σ(T )∗ for every T ∈ B[H] by Theorem 2.6, and
according to Proposition 2.Q(b),
)
σ(p(T, T ∗ )) = p(Φ(T ), Φ(T )) ∈ C : Φ ∈ A(T
)}
= p(λ, λ) ∈ C : λ ∈ {Φ(T ) ∈ C : Φ ∈ A(T
= p(λ, λ) ∈ C : λ ∈ σ(T )
= p(σ(T ), σ(T )∗ ) = p(σ(T ), σ(T ∗ )).
38 2. Spectrum
The first identity in the above expression defines the spectral radius rσ (T ),
and the second one is a consequence of the Weierstrass Theorem (cf. proof of
Theorem 2.2) since σ(T ) = ∅ is compact in C and the function | |: C → R is
continuous. A straightforward consequence of the Spectral Mapping Theorem
for polynomials reads as follows.
The next result is the well-known Gelfand–Beurling formula for the spec-
tral radius. Its proof requires another piece of elementary complex analysis,
viz., every analytic function has a power series representation. That is, if
f : Λ → C is analytic, and if Bα,β (ν) = {λ ∈ C : 0 ≤ α < |λ − ν| < β} lies in
the open set Λ ⊆ C , then f has a unique Laurent expansion about the point
∞
ν, namely, f (λ) = k=−∞ γk (λ − ν)k for every λ ∈ Bα,β (ν).
Proof. Since rσ (T )n ≤ T n for every positive integer n, and since the limit
1
of the sequence {T n n} exists by Lemma 1.10, we get
1
rσ (T ) ≤ lim T n n .
n
For the reverse inequality, proceed as follows. Consider the Neumann expan-
sion (Theorem 1.3) for the resolvent function RT : ρ(T ) → G[X ],
∞
RT (λ) = (λI − T )−1 = λ−1 T k λ−k
k=0
for every λ ∈ ρ(T ) such that T < |λ|, where the above series converges in
the (uniform) topology of B[X ]. Take an arbitrary bounded linear functional
∗
η : B[X ] → C in B[X ] (cf. proof of Theorem 2.2). Since η is continuous,
∞
η(RT (λ)) = λ−1 η(T k )λ−k
k=0
The converses to the above one-way implications fail in general. The next
result applies the preceding characterization of uniform stability to extend
the Neumann expansion of Theorem 1.3.
T k T k
Therefore, (λI − T )−1 = λ1 ∞ , where ∞ ∈ B[X ] is the strong
n T k k=0 λ k=0 λ
limit of k=0 λ , which concludes the proof of (b).
It can be shown that W (T ) is a convex set in C (see, e.g., [50, Problem 210]),
and it is clear that
W (T ∗ ) = W (T )∗ .
42 2. Spectrum
σP (T ) ∪ σR (T ) ⊆ W (T ).
σ(T ) = σR (T ) ∪ σAP (T ) ⊆ W (T )− .
Unlike the spectral radius, the numerical radius is a norm on B[H]. That is,
0 ≤ w(T ) for every T ∈ B[H] and 0 < w(T ) if T = O, w(αT ) = |α|w(T ), and
w(T + S) ≤ w(T ) + w(S) for every α ∈ C and every S, T ∈ B[H]. However, the
numerical radius does not have the operator norm property in the sense that
the inequality w(S T ) ≤ w(S)w(T ) is not true for all operators S, T ∈ B[H].
Nevertheless, the power inequality holds: w(T n ) ≤ w(T )n for all T ∈ B[H] and
every positive integer n (see, e.g., [50, p. 118 and Problem 221]). Moreover,
the numerical radius is a norm equivalent to the (induced uniform) operator
norm of B[H] and dominates the spectral radius, as in the next theorem.
1
|T x ; y| ≤ 4 |T (x + y) ; (x + y)| + |T (x − y) ; (x − y)|
+ |T (x + iy) ; (x + iy)| + |T (x − iy) ; (x − iy)|
≤ 14 w(T ) x + y2 + x − y2 + x + iy2 + x − iy2 .
whenever x = y = 1. Thus, since T = supx=y=1 |T x ; y| (see, e.g.,
[66, Corollary 5.71]), it follows that T ≤ 2w(T ).
The scalar 0 may be anywhere. That is, if T ∈ B∞[H], then λ = 0 may lie
in σP (T ), σR (T ), σC (T ), or ρ(T ). However, if T is a compact operator on a
nonzero space H and 0 ∈ ρ(T ), then H must be finite dimensional. Indeed,
if 0 ∈ ρ(T ), then T −1 ∈ B[H] so that I = T −1 T is compact (since B∞[H] is
an ideal of B[H]), which implies that H is finite dimensional (cf. Proposition
1.Y). The preceding theorem in fact is a rewriting of the Fredholm Alternative
(and it is also referred to as the Fredholm Alternative). It will be applied
often from now on. Here is a first application. Let B0 [H] denote the class of
all finite-rank operators on H (i.e., the class of all operators from B[H] with
a finite-dimensional range). Recall that B0 [H] ⊆ B∞[H] (finite-rank operators
are compact — cf. Proposition 1.X). Let #A denote the cardinality of a set
A, so that #A < ∞ means “A is a finite set”.
Proof. If dim H < ∞, then an injective operator is surjective, and linear mani-
folds are closed (see, e.g., [66, Problem 2.18 and Corollary 4.29]), and so the di-
agram of Section 2.2 says that σ(T ) = σP (T ) = σP4(T ) (for σP1(T ) = σR1(T ∗ )∗
according to Theorem 2.6). On the other hand, suppose dim H = ∞. Since
B0 [H] ⊆ B∞[H], Theorem 2.18 says that
σ(T )\{0} = σP (T )\{0} = σP4(T )\{0}.
Since dim R(T ) < ∞ and dim H = ∞, it follows that R(T )− = R(T ) = H and
N (T ) = {0} (because dim N (T ) + dim R(T ) = dim H; see, e.g., [66, Problem
2.17]). Then 0 ∈ σP4(T ) (cf. diagram of Section 2.2), and therefore
σ(T ) = σP (T ) = σP4(T ).
If σP (T ) is infinite, then there exists an infinite set of linearly independent
eigenvectors of T (Theorem 2.3). Since every eigenvector of T lies in R(T ),
this implies that dim R(T ) = ∞ (because every linearly independent subset
of a linear space is included in some Hamel bases — see, e.g., [66, Theorem
2.5]), which is a contradiction. Conclusion: σP (T ) must be finite.
Mn ⊂ Mn+1
n+1 n
(λn+1 I − T )yn+1 = αi (λn+1 − λi )xi = αi (λn+1 − λi )xi ∈ Mn .
i=1 i=1
Recall that λn = 0 for all n, take any pair of integers 1 ≤ m < n, and set
y = ym − λ−1 −1
m (λm I − T )ym + λn (λn I − T )yn
so that T (λ−1 −1
m ym ) − T (λn yn ) = y − yn . Since y lies in Mn−1 ,
1
2 < y − yn = T (λ−1 −1
m ym ) − T (λn yn ),
Proposition 2.A. Let H = {0} be a complex Hilbert space and let T denote
the unit circle about the origin of the complex plane.
(a) If H ∈ B[H] is hyponormal, then σP (H)∗ ⊆ σP (H ∗ ) and σR (H ∗ ) = ∅.
(b) If N ∈ B[H] is normal, then σP (N ∗ ) = σP (N )∗ and σR (N ) = ∅.
(c) If U ∈ B[H] is unitary, then σ(U ) ⊆ T .
(d) If A ∈ B[H] is self-adjoint, then σ(A) ⊂ R .
(e) If Q ∈ B[H] is nonnegative, then σ(Q) ⊂ [0, ∞).
(f) If R ∈ B[H] is strictly positive, then σ(R) ⊂ [α, ∞) for some α > 0.
(g) If E ∈ B[H] is a nontrivial projection, then σ(E) = σP (E) = {0, 1}.
48 2. Spectrum
Proposition 2.B. Similarity preserves the spectrum and its parts, and so it
preserves the spectral radius. That is, let H and K be nonzero complex Hilbert
spaces. For every T ∈ B[H] and W ∈ G[H, K],
(a) σP (T ) = σP (W T W −1 ),
(b) σR (T ) = σR (W T W −1 ),
(c) σC (T ) = σC (W T W −1 ).
Hence σ(T ) = σ(W T W −1 ), ρ(T ) = ρ(W T W −1 ), and r(T ) = r(W T W −1 ).
Unitary equivalence also preserves the norm: if W is a unitary transforma-
tion, then, in addition, T = W T W −1 .
Proposition 2.E. Take any operator T ∈ B[H]. Let d denote the usual dis-
tance in C . If λ ∈ ρ(T ), then
Proposition 2.I. If H and K are complex Hilbert spaces and T ∈ B[H], then
2.7 Additional Propositions 49
r(T ) = inf W T W −1 .
W ∈G[H,K]
Proposition 2.M. Let D and T = ∂ D denote the open unit disk and the unit
circle about the origin of the complex plane, respectively. If S+ ∈ B[K+] is a
unilateral shift and S ∈ B[K] is a bilateral shift on complex spaces, then
(a) σP (S+ ) = σR (S+∗ ) = ∅, σR (S+ ) = σP (S+∗ ) = D , σC (S+ ) = σC (S+∗ ) = T .
(b) σ(S) = σ(S ∗ ) = σC (S ∗ ) = σC (S) = T .
2
A unilateral weighted shift T+ = S+ D+ in B[+ (H)] isthe product of a
∞
canonical unilateral shift S+ and a diagonal operator D+ = k=0 αk I, both in
2
B[+ (H)], where {αk }∞
k=0 is a bounded sequence of scalars. A bilateral weighted
shift T = SD in B[ 2 (H)] is the product of a canonical bilateral shift S and
∞
a diagonal operator D = k=−∞ αk I, both in B[ 2 (H)], where {αk }∞ k=−∞ is
a bounded family of scalars.
2
Proposition 2.N. Let T+ ∈ B[+ (H)] be a unilateral weighted shift, and let
2
T ∈ B[ (H)] be a bilateral weighted shift, where H is complex .
(a) If αk → 0 as |k| → ∞, then T+ and T are compact and quasinilpotent .
If , in addition, αk = 0 for all k, then
(b) σ(T+) = σR (T+) = σR2(T+) = {0} and σ(T+∗ ) = σP (T+∗ ) = σP2(T+∗ ) = {0},
(c) σ(T ) = σC (T ) = σC (T ∗ ) = σ(T ∗ ) = {0}.
Proposition 2.Q. Let A be any unital complex Banach algebra (for instance,
A = B[X ]). If T ∈ A , where A is any closed unital subalgebra of A, then
(a) ρ (T ) ⊆ ρ(T ) and r (T ) = r(T ) (invariance of the spectral radius).
Hence ∂σ (T ) ⊆ ∂σ(T ) and σ(T ) ⊆ σ (T ).
Thus σ (T ) is obtained by adding to σ(T ) some holes of σ(T ).
(b) If A is a maximal commutative subalgebra of A, then
σ(T ) = Φ(T ) ∈ C : Φ ∈ A for each T ∈A.
This is the Rosenblum Corollary, which will be used to prove the Fuglede
Theorems in Chapter 3.
Notes: Again, as in the previous chapter, these are basic results that will be
needed throughout the text. We will not point out here the original sources
but, instead, well-known secondary sources where the reader can find the
proofs for those propositions, as well as some deeper discussions on them.
Proposition 2.A holds independently of the forthcoming Spectral Theorem
(Theorem 3.11) — see, e.g., [66, Corollary 6.18]. A partial converse, however,
needs the Spectral Theorem (see Proposition 3.D in Chapter 3). Proposition
2.B is a standard result (see, e.g., [50, Problem 75] and [66, Problem 6.10]), as
is Proposition 2.C (see, e.g., [50, Problem 76]). Proposition 2.D also dismisses
the Spectral Theorem of the next chapter; it is obtained by the Square Root
Theorem (Proposition 1.M) and by the Spectral Mapping Theorem (Theorem
2.7). Proposition 2.E is a rather useful technical result (see, e.g., [63, Problem
6.14]. Proposition 2.F is a synthesis of some sparse results (cf. [28, Proposition
I.5.1], [50, Solution 98], [55, Theorem 5.42], [66, Problem 6.37], and Propo-
sition 2.E). For Propositions 2.G and 2.H see [66, Sections 6.3 and 6.4]. The
spectral radius formula in Proposition 2.I (see, e.g., [42, p. 22]) ensures that an
operator is uniformly stable if and only if it is similar to a strict contraction
(hint: use the equivalence between (a) and (b) in Corollary 2.11). For Propo-
sitions 2.J and 2.K see [66, Sections 6.3 and 6.4]. The examples of spectra
in Propositions 2.L, 2.M, and 2.N are widely known (see, e.g., [66, Examples
6.C, 6.D, 6.E, 6.F, and 6.G]). Proposition 2.O deals with the equivalence be-
tween uniform and weak stabilities for compact operators, and it is extended
to a wider class of operators in Proposition 2.P (see [63, Problems 8.8. and
8.9]). Proposition 2.Q(a) is readily verified, where the invariance of the spec-
tral radius follows by the Gelfand–Beurling formula. Proposition 2.Q(b) is a
key result for proving both Theorem 2.8 (the Spectral Mapping Theorem for
normal operators) and also Lemma 5.43 of Chapter 5 (for the characterization
2.7 Additional Propositions 53
of the Browder spectrum in Theorem 5.44) — see, e.g., [76, Theorems 0.3 and
0.4]. For Propositions 2.R, 2.S, and 2.T see [80, Theorem 11.22], [76, Theorem
0.8], and [76, Corollary 0.13], respectively. Proposition 2.T will play a central
role in the proof of the Fuglede Theorems of the next chapter (precisely, in
Corollary 3.5 and Theorem 3.17).
Suggested Reading