Mat U
Mat U
Jan Dereziński
Contents
1 Unbounded operators                                                                                                                                                     2
  1.1 Relations . . . . . . . . . . . . . .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   2
  1.2 Linear pseudooperators . . . . . .      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   3
  1.3 Closed operators . . . . . . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   3
  1.4 Closable operators . . . . . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   4
  1.5 Perturbations of closed operators .     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   5
  1.6 Invertible unbounded operators . .      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   6
  1.7 Spectrum of unbounded operators         .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   7
  1.8 Examples of unbounded operators         .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   8
  1.9 Pseudoresolvents . . . . . . . . . .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   9
                                                              1
    3.10   Relative operator boundedness      .   .   .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   22
    3.11   Relative form boundedness . .      .   .   .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   24
    3.12   Non-maximal operators . . . .      .   .   .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   24
    3.13   Hermitian operators II . . . . .   .   .   .    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   25
4 Sesquilinear forms                                                                                                                                                               26
  4.1 Sesquilinear forms . . . . . . . . . . . . .                     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   26
  4.2 Closed positive forms . . . . . . . . . . . .                    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   27
  4.3 Closable positive forms . . . . . . . . . . .                    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   27
  4.4 Operators associated with positive forms .                       .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   28
  4.5 Polar decomposition . . . . . . . . . . . .                      .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   28
  4.6 Sectorial forms . . . . . . . . . . . . . . .                    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   29
  4.7 Operators associated with sectorial forms                        .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   30
  4.8 Perturbations of sectorial forms . . . . . .                     .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   30
  4.9 Friedrichs extensions . . . . . . . . . . . .                    .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   .   31
1     Unbounded operators
1.1    Relations
Let X, Y be sets. R is called a relation iff R ⊂ Y × X. We will also write R : X → Y . (Note the inversion
of the direction). An example of a relation is the identity 1X := {(x, x) : x ∈ X} ⊂ X × X.
    Introduce the “projections”
Y × X 3 (y, x) 7→ πY (y, x) := y ∈ Y,
                                    Y × X 3 (y, x) 7→ πX (y, x) := x ∈ X,
and the “flip”
                                Y × X 3 (y, x) 7→ τ (y, x) := (x, y) ∈ X × Y.
The domain of R is defined as DomR := πX R, its range is Ran R = πY R, the inverse of R is defined as
R−1 := τ R ⊂ X × Y . If S ⊂ Z × Y , then the superposition of S and R is defined as S ◦ R ⊂ Z × X,
S ◦ R := {(z, x) ∈ Z × X : ∨ (z, y) ∈ S, (y, x) ∈ R}.
                              y∈Y
    If X0 ⊂ X, then the restriction of R to X0 is defined as
                                              R                := R ∩ Y ×X0 .
                                                      X0
                                                                       2
If, moreover, Y0 ⊂ Y , then
                                          R            := R ∩ Y0 ×X0 .
                                              X0 →Y0
    A right unique relation will be also called a pseudo-transformation (-operator, etc). Instead of writing
(y, x) ∈ R, we will then write y = R(x) or, in some contexts, y = Rx. We also introduce the graph of R:
Note that strictly speaking Gr R = R. The difference of Gr R and R lies only in their syntactic role.
   Note that a superposition of pseudotransformations is a pseudotransformation.
   We say that a pseudotransformation is injective if it is left-unique. The inverse of a pseudotransfor-
mation is a pseudotransformation iff it is injective.
   A transformation is a pseudotransformation with a maximal domain. The composition of transfor-
mations is a transformation.
   We say that a transformation R is bijective iff it is left-unique and Ran R = Y . The inverse of a
transformation is a transformation iff it is bijective.
Proposition 1.2 Let R ⊂ X ×Y and S ⊂ Y ×X be transformations such that R◦S = 1Y and S◦R = 1X .
Then S and R are bijections and S = R−1 .
   From now on by an “operator” we will mean a “pseudooperator”. To say that A is a true operator
we will write DomA = X .
is a Banach space.
                                                         3
Proof. The equivalence of (1), (2) and (3) is obvious, if we note that
DomA 3 x 7→ (Ax, x) ∈ Gr A
is a bijection. 2
B(X , Y) will denote all bounded operators from X to Y with the domain equal to X .
Theorem 1.7 Let B ∈ B(X , Y) be invertible and A : X → Y closed. Then BA is closed on DomA and
AB is closed on B −1 DomA.
Definition 1.10 An operator A satisfying the conditions of Theorem 1.9 is called closable. If the con-
ditions of Theorem 1.9 hold, then the operator whose graph equals (Gr A)cl is denoted by Acl and called
the closure of A.
    Proof of Theorem 1.9 To show (2)⇒(1) it suffices to take as B the operator Acl . Let us show
(1)⇒(2). Let B be a closed operator such that A ⊂ B. Then (Gr A)cl ⊂ (Gr B)cl = Gr B. But
(y, 0) ∈ Gr B ⇒ y = 0, hence (y, 0) ∈ (Gr A)cl ⇒ y = 0. Thus (Gr A)cl is the graph of an
operator. 2
    As a by-product of the above proof, we obtain
                                                    4
Proposition 1.11 If A is closable, B closed and A ⊂ B, then Acl ⊂ B.
Proposition 1.12 Let A be bounded. Then A is closable, DomAcl = (DomA)cl and Acl x = limn Axn
for xn → x, xn ∈ DomA. Besides, Acl satisfies (1.1).
      n→∞
   Let A be a closed operator. We say that a linear subspace D is an essential domain for A iff D is
dense in DomA in the graph topology. In other words, D is an essential domain for A, if
                                                         cl
                                                  A               = A.
                                                      D
Theorem 1.13 (1) If A ∈ B(X , Y), then a linear subspace D ⊂ X is an essential domain for A iff it
   is dense in X (in the usual topology).
(2) If A is closed, has a dense domain and D is its essential domain, then D is dense in X .
The infimum of a satisfying (1.2) is called the A-bound of B. In other words: the A-bound of B equals
                                                                    kBxk
                                       inf    sup                            .
                                      c>0 x∈DomA\{0}             kAxk + ckxk
Theorem 1.15 B is bounded relatively to A with the A-bound a iff DomA ⊂ DomB and
                                                                           1/2
                                                       kBxk2
                                                
                                inf    sup                                        < ∞.                 (1.3)
                               c>0 x∈DomA\{0}       kAxk2 + ckxk2
Theorem 1.16 Let A be closed and let B be bounded relatively to A with the A-bound less than 1. Then
A + B with the domain DomA is closed. All essential domains of A are essential domains of A + B
                                                             5
Proof. We know that
                                               kBxk ≤ akAxk + bkxk
for some a < 1 and b. Hence
and
              (1 − a)kAxk + kxk ≤ kAxk − kBxk + (1 + b)kxk ≤ k(A + B)xk + (1 + b)kxk.
Hence the norms kAxk + kxk and k(A + B)xk + kxk are equivalent on DomA. 2
Theorem 1.17 Suppose that A, C are two operators with the same domain DomA = DomC = D
satisfying
                           k(A − C)xk ≤ a(kAxk + kCxk) + bkxk
for some a < 1. Then
 (1) A is closed on D iff C is closed on D.
(2) A is closable on D iff C is closable on D and then the domains of Acl and C cl coincide.
Hence
                                                   2a             b
                                        kBxk ≤        kF (t)xk +     kxk.
                                                  1−a            1−a
Therefore, if |s| <   1−a
                       2a   and t, t + s ∈ [0, 1], then F (t + s) is closed iff F (t) is closed. 2
kA−1 k ≤ c−1
Proof. Let yn ∈ Ran A and yn → y. Let Axn = yn . Then xn is a Cauchy sequence. Hence there exists
limn→∞ xn := x. But A is closed, hence Ax = y. Therefore, Ran A is closed. 2
kAxk ≥ ckxk,
                                                           6
Theorem 1.21 Let A be closable and for some c (1.4) holds. Then (1.4) holds for Acl as well.
Theorem 1.22 (1) Let A be injective and DomB ⊃ DomA. Then B has the A-bound less than
   kBA−1 k.
(2) If, moreover, kBA−1 k < 1, then A + B with the domain DomA is closed, invertible and
                                                   ∞
                                                   X
                                     (A + B)−1 =         (−1)j A−1 (BA−1 )j .
                                                   j=0
we see that B has the A-bound less than or equal to a. This proves (1).
   Assume now that a < 1. Let
                                          X n
                                    Cn :=     (−1)j A−1 (BA−1 )j .
                                             j=0
B −1 − A−1 = B −1 (A − B)A−1 .
rsA := {z ∈ C : z − A is invertible }.
                                                    7
Theorem 1.24 (1) If λ, µ ∈ rsA, then
Proposition 1.25 Suppose that rsA is non-empty and DomA is dense. Then DomA2 is dense.
Proof. Let z ∈ rsA. (z − A)−1 is a bounded operator with a dense range and DomA is dense. Hence
(z − A)−1 DomA is dense. If x ∈ DomA, then A(z − A)−1 x = (z − A)−1 Ax ∈ DomA Hence (z −
A)−1 DomA ⊂ DomA2 . 2
Theorem 1.26 Let A and B be operators on X with A ⊂ B, A 6= B. Then rsA ⊂ spB, and hence
rsB ⊂ spA.
Proof. Let λ ∈ rsA. Let x ∈ DomB\DomA. We have Ran (λ − A) = X , hence there exists y ∈ DomA
                                                                  6 rsB. 2
such that (λ − A)y = (λ − B)x. Hence (λ − B)y = (λ − B)x. Hence λ ∈
by the formula
                                                  (Ax)i = ai xi .
                                    p
(We can use C∞ (I) instead of L (I), then p = ∞ in the formulas below). Then the operator A is
unbounded and non-closed. Besides,
                                                        8
   The closure of A has the domain
We then have
                                         spp (Acl ) = {ai : i ∈ I},
                                         spAcl = {ai : i ∈ I}cl .
To prove this let D be the rhs of (1.5) and x ∈ D. Then there exists a countable set I1 such that i 6∈ I1
implies xi = 0. We enumerate the elements of I1 : i1 , i2 , . . .. Define xn ∈ C0 (I) setting xnij = xij
for j ≤ n and xni = 0 for the remaining indices. Then limn→∞ xn = x and Axn → Ax. Hence,
{(x, Ax) : x ∈ D} ⊂ (Gr A)cl .
    If xn belongs to (1.5) and (xn , Axn ) → (x, y), then xni → xi and ai xni = (Axn )i → yi . Hence
yi = ai xi . Using that y ∈ Lp (I) we see that x belongs to (1.5).
Example 1.28 Let p−1 + q −1 = 1, 1 < p ≤ ∞ and let (wi )i∈I be a sequence that does not belong to
Lq (I). Let C0 (I) be as above. Define
                                                               X
                               Lp (I) ⊃ C0 (I) 3 x 7→ hw|xi :=   xi wi ∈ C.
                                                                 i∈I
1.9    Pseudoresolvents
Definition 1.29 Let Ω ⊂ C be open. Then the continuous function
Ω 3 z 7→ R(z) ∈ B(X )
is called a pseudoresolvent if
                                  R(z1 ) − R(z2 ) = (z2 − z1 )R(z1 )R(z2 ).                          (1.6)
Proposition 1.30 Let Ω 3 z 7→ Rn (z) ∈ B(X ) be a sequence of pseudoresolvents and R(z) := s− limn→∞ Rn (z).
Then R(z) is a pseudoresolvent.
Proof. Let us prove (4)⇐. Fix z1 ∈ Ω. If N = {0}, then every element of R can be uniquely represented
as R(z1 )x, x ∈ X . Define HR(z1 )x := −x + z1 R(z1 )x. By formula (1.6) we check that the definition does
not depend on z1 . 2
                                                     9
2     One-parameter semigroups in Banach spaces
2.1    (M, β)-type semigroups
Let X be a Banach space.
Definition 2.1 [0, ∞[3 t 7→ W (t) ∈ B(X ) is called a strongly continuous one-parameter semigroup iff
(1) W (0) = 1;
(2) W (t1 )W (t2 ) = W (t1 + t2 ), t1 , t2 ∈ [0, ∞[;
(3) limt↓0 W (t)x = x, x ∈ X ;
(4) for some t0 > 0, kW (t)k < M , 0 ≤ t ≤ t0 .
Remark 2.2 Using the Banach-Steinhaus Theorem one can show that (4) follows from the remaining
points.
Proof. By (4), for t ≤ nt0 we have kW (t)k ≤ M n . Hence, kW (t)k ≤ M exp( tt0 log M ). Therefore, (2.7)
is satisfied.
    Let tn → t and xn → x. Then
                                                        10
    By (3), if A is an operator, which is a generator of a semigroup W (t), then such W (t) is unique. We
will write W (t) =: etA .
Proof of Theorem 2.4 (2). Let x ∈ DomA. Then
                    lim s−1 (W (s) − 1)W (t)x = W (t) lim s−1 (W (s) − 1)x = W (t)Ax.               (2.8)
                    s↓0                                       s↓0
Hence the limit of the left hand side of (2.8) exists. Hence W (t)x ∈ DomA and AW (t)x = W (t)Ax. 2
(2) is obvious. (4) is proven as Theorem 2.4 (2). To prove (3) we note that
2
Proof of Theorem 2.4 (1), (3) The density of DomA follows by Lemma 2.5 (1) and (3).
   Let us show that A is closed. Let xn → x and Axn → y. Using the boundedness of Bt A = ABt
                                        n→∞            n→∞
we get
                                     Bt y = lim Bt Axn = Bt Ax.
                                               n→∞
Hence
                                     y = lim Bt y = lim Bt Ax = Ax.
                                         t↓0              t↓0
2
Proposition 2.6 Let W (t) be a semigroup and A its generator. Then, for any x ∈ DomA there exists a
unique solution of
                                                             d
                              [0, t0 ] 3 t 7→ x(t) ∈ DomA,      x(t) = Ax(t),               (2.10)
                                                             dt
(for t = 0 the derivative is right-sided). The solution is given by x(t) = W (t)x.
Proof. Let us show that x(t) := W (t)x solves (2.10). We already know that the right-sided derivative
equals Ax(t). It suffices to prove the same about the left-sided derivative. For 0 ≤ u ≤ t we have
                                                          11
   Let us show now the uniqueness. Let x(t) solve (2.10) . Let y(s) := W (t − s)x(s). Then
                              d
                                 y(s) = W (t − s)Ax(s) − AW (t − s)x(s) = 0
                              ds
Hence y(s) does not depend on s. At s = t it equals x(t), and at s = 0 it equals W (t)x. 2
Proof of Theorem 2.4 (3) By Theorem 2.6 (2), W (t) is uniquely determined by A on DomA. But
W (t) is bounded and DomA is dense, hence W (t) is uniquely determined. 2
kBt − 1k < 1.
                                                     12
Hence Bt x ∈ D̃. By Lemma 2.9 (2)
                                           kBt x − xkDomA → 0.
                                                           t↓0
Hence, x ∈ D̃. 2
Proof. It suffices to prove (1)⇒(2). Let (1) be satisfied. It suffices to assume that β = 0. Let z = x + iy.
Then for t > 0
                        (z − A)−m = (x + t − A)m (1 + (iy − t)(x + t − A)−1 )−m
                                        ∞
                                                                              
                                       P              −m−j           j    −m
                                     =     (x + t − A)      (iy − t)             .
                                       j=0                                  j
                           
                       −m
Using the fact that            has an alternating sign we get
                         j
                                                                                  
                                −m
                                         P∞           −m−j        j        j   −m
                       k(z − A) k ≤ j=0 |x + t|             (−1) |iy − t|
                                                                                j
                                                               −m
                                       = M |x + t|m 1 − |iy−t|
                                                           x+t
Definition 2.11 We say that an operator A is (M, β)-type, iff the conditions of Theorem 2.10 are sat-
isfied.
                                                    13
Proof. Set                                                       Z     ∞
                                                 R(z)x :=                  e−zt W (t)xdt.
                                                                   0
Theorem 2.13 If A is an operator of (M, β)-type, then it is the generator of a semigroup W (t). This
semigroup is of (M, β)-type.
   To simplify, let us assume that β = 0 (which does not restrict the generality). Then we have the
formula
                                                          −n
                                                        t
                                   W (t) = s− lim 1 − A         ,
                                             n→∞        n
                                          −n
                                                     t2
                                 
                                       t
                         W (t)x − 1 − A       x ≤ M kA2 xk, x ∈ DomA2 .
                                      n              2
                                                                       14
   Let us list some other properties of Vn (t): for Ret > 0, Vn (t) is holomorphic, kVn (t)k ≤ M and
                                                            −n−1
                                      d                   t
                                         Vn (t) = A 1 − A             .
                                      dt                 n
                                                                     1 t2
                                                   = M 2 ( n1 +      m) 2 .
By the Proposition 1.25, Dom(A2 ) is dense in X . Therefore, there exists a limit uniform on [0, t0 ]
                                                     15
(5) {z ∈ C : Rez > 0} ⊂ rsA; besides, if x ∈ DomA, v ∈ X # , kvk = 1 and hv|xi = kxk, then
Rehv|Axi ≤ 0.
(6) There exists z ∈ C with Rez > 0 such that Ran (z − A) = X; besides if x ∈ DomA, then there
    ∨ v ∈ X # such that kvk = 1, hv|xi = kxk, and
Rehv|Axi ≤ 0.
Proof. The equivalence of (1) and (2) is a special case of Theorems 2.12 and 2.13. The implications
(2)⇒(3) and (2)⇒(4) are obvious, the converse implications are easy.
   Let us show (1),(3)⇒(5). We have
Hence
                              Rehv|Axi = lim Ret−1 (hv|etA xi − hv|xi) ≤ 0.
                                            t↓0
Using Ran (z − A)−1 = X , we conclude that (z − A)−1 exists and k(z − A)−1 k ≤ |Rez|−1 . 2
DomA 3 v 7→ (w|Av)
is bounded (in the topology of V). Hence there exists a unique y ∈ V such that
(w|Av) = (y|v), v ∈ V.
We set then
                                                  A∗ w = y.
                                                      16
   If CV , CW are the Riesz antiisomorphisms, then we have the relationship between the (Banach space)-
conjugate A# and the (Hilbert space)-adjoint A∗ :
A∗ = CV−1 A# CW .
Gr A∗ = j(Gr A)⊥ .
⇔ (w|Av) = 0, v ∈ DomA
                                             ⇔ w ∈ (Ran A)⊥ .
   Proof of (4)
                              v ∈ KerA ⇔ (w|Av) = 0, w ∈ W
⇒ (w|Av) = 0, w ∈ DomA∗
⇔ v ∈ (Ran A∗ )⊥ .
                                                  17
  (4) is proven in Theorem 3.2.
  To prove (5) note that in the second line of the proof of Theorem 3.2 (4) we can use the fact that
DomA∗ is dense in W to replace ⇒ with ⇔. 2
Hence
                          Gr A−1∗ = j(τ (Gr A))⊥ = τ −1 (j(Gr A)⊥ ) = Gr A∗−1 .
2
Theorem 3.5 Let A : V → W be densely defined and closed. Then the following conditions are equiva-
lent:
 (1) A is invertible.
(2) A∗ is invertible.
(3) For some c > 0, kAvk ≥ ckvk, v ∈ V and kA∗ wk ≥ ckvk, w ∈ W.
Moreover, spext (A) = spext (A∗ ).
Proof. (1)⇒(2). Let A be invertible. Then A−1 ∈ B(W, V). Hence, A−1∗ ∈ B(V, W).
   Clearly, the assumptions of Theorem 3.4 are satisfied, and hence A∗−1 = A−1∗ . Therefore, A∗−1 ∈
B(V, W).
   (1)⇐(2). A∗ is also densely defined and closed. Hence the same arguments as above apply.
   It is obvious that (1) and (2) imply (3). Let us prove that (3)⇒(1). kA∗ vk ≥ ckvk implies that
KerA∗ = {0}. Hence (Ran A)⊥ is dense. This together with kAvk ≥ ckvk implies that Ran A = W, and
consequently, A is invertible. 2
                                                     18
Proof. To prove (1), take (z0 6∈ NumT )cl . Recall that NumT is convex. Hence, replacing T wih αT + β
we can assume that z0 = iν and 0 ∈ NumT ⊂ {Imz ≤ 0}. Thus ν = dist(iν, NumT ) and
≥ |ν|2 kvk2 .
Theorem 3.8 Suppose that T is an operator and for any connected component ∆i of C\(NumT )cl we
choose λi ∈ ∆i . Then the following conditions are necessary and sufficient for T to be maximal
 (1) For all i, λi 6∈ spT ;
(2) T is closable and for all i, Ran (λi − T ) = V.
(3) T is closed and for all i, Ran (λi − T ) is dense in V.
(4) T is closed and for all i, Ker(λi − T ∗ ) = {0}.
Im(v|Av) ≤ 0, v ∈ DomA.
Theorem 3.9 Let A be a densely defined operator. Then the following conditions are equivalent:
(1) −iA is the generator of a strongly continuous semigroup of contractions.
(2) A is maximally dissipative.
Hence A is dissipative.
   We know that the generators of contractions satisfy {Rez > 0} ⊂ rs(−iA).
                                                       19
   (2)⇒(1) Let Rez > 0. We have
                           kvkk(z + iA)vk     ≥ |(v|(z + iA)v)|
Theorem 3.10 Let A be dissipative. Then the following conditions are equivalent:
(1) A is maximally dissipative.
(2) A is closable and there exists z0 with Imz0 > 0 and Ran (z0 − A) = V.
(3) A is closed and there exists z0 with Imz0 > 0 and Ran (z0 − A) dense in V.
(4) A is closed and there exists z0 with Imz0 > 0 and Ker(z 0 − A∗ ) = {0}.
Theorem 3.13 Let A be a densely defined operator. Then the following conditions are equivalent:
(1) −iA is the generator of a strongly continuous semigroup of isometries.
(2) A is hermitian and spA ⊂ {Imz ≤ 0}.
Proof. (1)⇒(2) For v ∈ DomA,
   (2)⇒(1) We know that e−itA is the generator of a strongly continuous contractive semigroup. For
v ∈ DomA,
                                      0 = ∂t (e−itA v|e−itA v)
Hence, for v ∈ DomA, ke−itA vk2 = kvk2 . 2
Theorem 3.14 Let A be hermitian. Then the following conditions are equivalent:
(1) spA ⊂ {Imz ≤ 0}.
(2) There exists z0 with Imz0 > 0 and Ran (z0 − A) = V.
(3) A is closed and there esxists z0 with Imz0 > 0 and Ran (z0 − A) dense in V.
(4) A is closed and there exists z0 with Imz0 > 0 and Ker(z 0 − A∗ ) = {0}.
                                                      20
3.7    Self-adjoint operators
Let T be a densely defined operator on V. T is self-adjoint iff T ∗ = T , that means if for w ∈ W there
exists y ∈ V such that
                                     (y|v) = (w|T v), v ∈ DomT,
then w ∈ DomT and T w = y.
Theorem 3.15 Every self-adjoint operator is hermitian and closed. If A is bounded, then it is self-
adjoint iff it is hermitian.
Theorem 3.16 Fix z± with ±Imz± > 0. Let A be hermitian. Then the following conditions are
necessary and sufficient for A to be self-adjoint:
 (1) spA ⊂ R.
(2) z± 6∈ spA.
(3) Ran (z± − A) = V.
(4) A is closed and Ran (z± − A) is dense in V.
(5) A is closed and Ker(z ± − A∗ ) = {0}.
Theorem 3.17 Let λ0 ∈ R. Let A be hermitian. Then the following conditions are sufficient for A to
be self-adjoint:
 (1) λ0 6∈ spA.
(2) Ran (λ0 − A) = V.
(3) A is closed and Ran (λ0 − A) is dense in V.
(4) A is closed and Ker(λ0 − A∗ ) = {0}.
Theorem 3.18 Let A be self-adjoint. Then U := (A + i)(A − i)−1 is a unitary operator with spU =
(spext A + i)(spext A − i)−1 .
Of course, we can also apply the functional calculus for measurable functions. In particular, the function
spA 3 x 7→ id(x) := x is a measurable function on spA. We have idA = A.
Theorem 3.19 (Stone Theorem) Let A be an operator. Then the following conditions are equiva-
lent:
 (1) iA is the generator of a strongly continuous group of unitary operators.
(2) A is self-adjoint.
Proof. To prove (1)⇒(2), suppose that R 7→ U (t) is a strongly continuous unitary group. Let −iA be
its generator. Then [0, ∞[3 U (t), U (−t) are semigroups of contractions with the generators iA and −iA.
By Theorem 3.19, A is hermitian and spA ⊂ R. Hence A is self-adjoint.
    (2)⇒(1) follows by the spectral theorem. 2
                                                    21
3.8    Essentially self-adjoint operators
Definition 3.20 An operator A : V → V is essentially self-adjoint iff Acl is self-adjoint.
Theorem 3.22 Fix z± with ±Imz± > 0 Let A be hermitian. Then the following conditions are necessary
and sufficient for A to be essentially self-adjoint:
 (1) A∗ is self-adjoint.
(2) Ran (z± − A) is dense in V.
(3) Ker(z ± − A∗ ) = {0}.
Theorem 3.23 Let λ0 ∈ R. Let A be hermitian. Then the following conditions are sufficient for A to
be self-adjoint:
 (1) Ran (λ0 − A) is dense in V.
(2) Ker(λ0 − A∗ ) = {0}.
and V−α := Vα∗ , (where Vα∗ denotes the space of bounded antilinear functionals on Vα ). Note that we
have the identification V = V ∗ , hence both definitions give V0 = V.
   It is clear, that for 0 ≤ α ≤ β, V ⊃ Vα ⊃ Vβ . Hence V−α = Vα∗ can be identified with a subspace of
V−β = Vβ∗ . Thus we obtain Vα ⊃ Vβ for any α ≤ β.
   Note that for α > 0 V is embedded in V−α and for v, w ∈ V
                                                                  
                                      (v|w)−α = B −α/2 v|B −α/2 w .
                                                    22
(1) B has the A-bound equal to a1 , that is
                                                                                21
                                                                kBvk2
                                                        
                                    inf      sup                                      = a1 .
                                   ν>0 v6=0, v∈DomA          kAvk2 + ν 2 kvk
(2)
Theorem 3.26 (Kato-Rellich) Let A be self-adjoint, B hermitian. Let B be A-bounded with the
A−bound < 1. Then
 (1) A + B is self-adjoint on DomA.
(2) If A is essentally self-adjoint on D, then A + B is essentially self-adjoint on D.
Proof. Clearly, A + B is hermitian on DomA. Moreover, for some ν, kB(iν − A)−1 k < 1 and (which
is equivalent by the unitarity of (A − iν)(A + iν)−1 ), kB(−iν − A)−1 k < 1. Hence, iν − A − B and
−iν − A − B are invertible. 2
      Let us note an improved version of the notion of the operator boundedness:
Theorem 3.27 Let A be a closed operator and B an operator with DomB ⊃ DomA. Then the following
statements are equivalent:
 (1)
                                                                                       12
                                                                  kBvk2
                                                        
                                    a2 = inf sup                                             .
                                           µ,ν>0 v6=0       k(A − µ)vk2 + ν 2 kvk
(2)
Note hat the analog of Theorem 3.26 is true with a1 replaced with a2 .
                                                            23
3.11        Relative form boundedness
Theorem 3.28 Let A be a self-adjoint operator. Let B be a bounded operator from (1 + |A|)−1/2 H to
(1 + |A|)1/2 H. Then the following statements are equivalent:
 (1)
                                                          1                      1
                                inf k(A − µ)2 + ν 2 )− 4 B((A − µ)2 + ν 2 )− 4 k = a3 .
                                µ,ν
 (2)
                                                              1              1
                                   inf k(µ + iν − A)− 2 v, (µ + iν − A)− 2 k = a3 .
                                  µ,ν>0
       If the conditions of the above theorem are satified, then we say that the A-form-bound of B equals
a3 .
Theorem 3.29 Let A be a self-adjoint operator. Let B have the A-form-bound less than 1. Then there
exists a open subsets in the upper and lower complex half-plane such that the series
                                                ∞
                                                X
                                      R(z) :=         (z − A)−1 (B(z − A)−1 )j
                                                j=0
is convergent. Moreover, R(z) is a resolvent of a self-adjoint operator, which will be called the form sum
                                                                              1             1
of A and B. If A is bounded from below, then so is A + B and Dom|A + B| 2 = Dom|A| 2 .
Let w ∈ Ran (iλ − A). Then there exists v ∈ DomA such that
w = (iλ − A)v
and kvk ≤ λ−1 kwk. If moreover, w ∈ Ran (iλ + α − A)⊥ = Ker(−iλ − α − A∗ ), then
                                          0 = ((−iλ + α − A∗ )w|v)
                                          = (w|(iλ − A)v) + α(w|v)
                                          = kwk2 + α(w|v).
But
                                  |kwk2 + α(w|v)| ≥ (1 − |α|/|λ|)kwk2 > 0,
which is a contradiction and completes the proof of (3.14).
   Now (3.14) implies that dim Ran (iλ − A) ≤ dim Ran (iλ + α − A). 2
                                                           24
3.13    Hermitian operators II
Let A be closed hermitian.
Theorem 3.31 The so-called defect indices
n± := dim Ker(z − A∗ ), z ∈ C±
                      1) spA ⊂ R,            n± = 0, A is self-adjoint;
                      2) spA = {Imz ≥ 0},    n+ 6= 0, n− = 0, A is not self-adjoint;
                      3) spA = {Imz ≤ 0},    n+ = 0, n− 6= 0, A is not self-adjoint;
                      4) spA = C             n+ 6= 0, n− 6= 0, A is not self-adjoint.
Definition 3.32 Let A be hermitian and closed. Define on DomA∗ the following scalar product:
The A∗ −completeness and the A∗ −orthogonality is defined using the scalar product (·|·)A∗ . A space is
A∗ −hermitian iff [·|·]A∗ vanishes on this subspace.
and the components in the above direct sum are A∗ -closed, A∗ −orthogonal and
Proof. (1) is obvious. In (2) the A∗ −orthogonality and the A∗ −closedness are easy.
  Let w ∈ DomA∗ and
                                w ⊥ DomA ⊕ Ker(A∗ + i) ⊕ Ker(A∗ − i)
in the sense of the product (·|·)A∗ . In particular, for v ∈ DomA we have
                                                    25
Thus
                                         (A∗ − i)w ∈ Ker(A∗ + i).                                     (3.15)
     If y ∈ Ker(A∗ + i), then
w ∈ Ker(A∗ − i).
If
                                               A ⊂ B ⊂ A∗ ,
then the subspace Z corresponding to B will be denoted by ZB . We will write
Z := {z ⊕ U z : z ∈ Ran U ∗ U }.
4      Sesquilinear forms
4.1     Sesquilinear forms
Let V, W be complex spaces. We say that t is a sesquilinear quasiform on W × V iff there exist subspaces
Doml t ⊂ W and Domr t ⊂ V such that
is a sesquilinear map. From now on by a sesquilinear form we will mean a sesquilinear quasiform.
    We define a form t∗ with the domains Doml t∗ := Domr t, Domr t∗ := Doml t, by the formula t∗ (v, w) :=
t(w, v). If t1 are t2 forms, then we define t1 + t2 with the domain Doml t1 + t2 := Doml t1 ∩ Doml t1 ,
Domr t1 + t2 := Domr t1 ∩ Domr t1 by (t1 + t2 )(w, v) := t1 (w, v) + t2 (w, v). We write t1 ⊂ t2 if Doml t1 ⊂
Doml t2 , Domr t1 ⊂ Domr t2 , and t1 (w, v) = t2 (w, v) w ∈ Doml t1 , v ∈ Domr t1 .
    t is bounded iff
                               |t(w, v)| ≤ ckwkkvk, w ∈ Doml t, v ∈ Domr t.
   From now on, we will usually assume that W = V and Doml t = Domr t and the latter subspace will
be simply denoted by Domt.
                                                     26
   Recall that the numerical range of the form t is defined as
                                 Numt := {t(v) : v ∈ Domt, kvk = 1}.
We proved that Numt is a convex set.
  The form t is bounded iff Numt is bounded. Equivalently, |t(v)| ≤ ckvk2 .
  t is bounded from below, if there exists c such that
                                        Numt ⊂ {z : Rez > c}.
   t is hermitian iff Numt ⊂ R. The equivalent condition: t(w, v) = t(v, w).
   If T is an operator on V, then t(w, v) := (w, T v) with the domain DomT is a form called the form
associated with the operator T . Clearly, Numt = NumT .
Example 4.6 Let A be an operator. Then (Aw|Av) with the domain DomA is closable iff A is a closable
operator. Then (Acl w|Acl v) with the domain DomAcl is its closure.
                                                   27
4.4    Operators associated with positive forms
Let S be a positive self-adjoint operator. We define the form s as follows: Doms := DomS 1/2 and
s(v, w) := (S 1/2 v|S 1/2 w). We will say that s is the form associated with S.
Theorem 4.8 Let s be a densely defined closed positive form. Then there exists a unique positive self-
adjoint operator S such that Doms = DomS 1/2 and s(v, w) := (S 1/2 v|S 1/2 w). We will say that S is the
operator associated with the form s.
    Let us show that Doms is an essential domain for s. Let v ∈ Doms is s-orthogonal to Ran A = DomS.
Then v is orthogonal to Doms—see (4.16). Hence v = 0.
                                 1                   1
    Define s1 by Doms1 = DomS 2 and s1 (w, v) = (S 2 w|S 1/2 v). The form s and s1 coincide on DomS ⊂
       1
DomS 2 ∩ Doms. We proved above that DomS is an essential domain for s. The form s1 is obviously
closed and Doms1 is an essential domain for s1 . Hence, s1 = s. 2
                                                      28
Proof. (1) is obvious.
  To see (2), note that for v ∈ DomA∗ , w ∈ DomA we have
Hence kA∗ vk−1 ≤ kvk, and so A∗ : V → V−1 is bounded. DomA∗ is dense in V. Hence A∗ extends to a
bounded operator A∗0 : V → V−1 .
   To prove (3), let v ∈ V, A∗0 v = w ∈ V. Then there exists (vn ) ⊂ DomA∗ such that vn → v in the
norm of V and A∗ vn → w in the norm of V−1 . Hence for x ∈ DomA,
Now
                                      DomB     = {v ∈ V1 : B1 v ∈ V}
= {v ∈ V1 : A∗0 A1 v ∈ V}
                                               = {v1 : Av ∈ DomA∗ }.
2
    Motivated by the above theorem we will write A∗ A for B.
Theorem 4.10 Let A be closed. Then there exist a unique positive operator |A| and a unique partial
isometry U such that KerU = KerA and A = U |A|. We have then Ran U = Ran Acl .
Proof. The operator A∗ A is positive. By the spectral theorem, we can then define
                                                  √
                                           |A| := A∗ A.
Lemma 4.11 Let t be a sectorial form with the sector given by a, θ. Then
                                                                      1              1
                           |(t − a)(w, v)| ≤ (1 + tan θ)Re(t − a)(w) 2 Re(t − a)(v) 2 .
                                                       29
    Clearly, if t is sectorial and a is the tip of a sector containing Numt, then Ret + Rea is a positive form.
We say that a sectorial form t is closed iff Ret is closed. We say that a sectorial form t is closable iff Ret
is closable.
    It is easy to see, using Lemma 4.11, that Theorems 4.2 and 4.5 remain true if the form s is assumed
to be sectorial.
Theorem 4.13 (1) Let t be a sectorial form. Then there exists a unique m-sectorial operator T such
   that DomT ⊂ Domt and
We will assume that the sector of t has the tip at 0. We will write s := Ret.
Proof. By Lemma 4.11, the form t + 1 is bounded in the Hilbert space Doms. Hence there exists
B ∈ B(Doms) such that
                        (t + 1)(w, v) = (w|Bv)s = (w|Bv) + s(w, Bv).
We have
                              kvk2s = Re(t + 1)(v) = Re(v|Bv)s ≤ kBvks kvks .
Hence kvks ≤ kBvks . Therefore, Ran B is closed.
   If w is orthogonal in Doms to Ran B, then
kwk2s = Re(w|Bw)s = 0.
                                                      30
Definition 4.16 Let p, s be forms and let s be positive. We say that p is s-bounded iff Doms ⊂ Domp
and
                                                     |p(v)|
                                 b := inf sup                 < ∞.
                                      c>0 v∈Doms s(v) + ckvk2
Theorem 4.17 Let t be sectorial and let p be Ret-bounded with the Ret-bound < 1. Then
(1) The form t + p (with the domain Domt) is sectorial as well.
(2) t is closed ⇔ t + p is closed.
(3) t is closable ⇔ t + p is closable, and then Dom(t + p)cl = Domtcl .
Hence
                      Im(t + p)(v) ≤ |Imt(v)| + |Imp(v)| ≤ (tan θ + b)Ret(v) + ckvk2 ,
                        Re(t + p)(v) ≥ Ret(v) − |Imp(v)| ≥ (1 − b)Ret(v) − ckvk2 .                  (4.17)
Hence,
                  |Im(t + p)(v)| ≤ (1 − b)−1 (tan θ + b) Re(t + p)(v) + ckvk2 + ckvk2 .
                                                                             
Proof. It suffices to assume that the tip of the sector of t is 0. Suppose that wn ∈ DomT = Domt,
wn → 0, limn,m→∞ t(wn − wm ) = 0. Then
For any  > 0 there exists N such that for n, m > N we have Ret(wn −wm ) ≤ 2 . Besides, limm→∞ (wm |T wn ) =
0. Therefore,
                                     |t(wn )| ≤ (1 + tan θ)|t(wn )|1/2 .
Hence t(wn ) → 0. 2
    Thus there exists a unique m-sectorial operator TFr associated with the form tcl . The operator TFr is
called the Friedrichs extension of T .
                                                       31
5     Aronszajn-Donoghue and Friedrichs Hamiltonian and their
      renormalization
5.1    Aronszajn Donoghue Hamiltonians
Let H0 be a self-adjoint operator on H, h ∈ H and λ ∈ R.
H := H0 + λ|h)(h|, (5.18)
is a rank one perturbation of H0 . We will call (5.18) the Aronszajn Donoghue Hamiltonian.
    We would like to describe how to define the Aronszajn-Donoghue Hamiltonian if h is not necessarily
a bounded functional on H. It will turn out that it is natural to consider 3 types of h:
where by H−n we denoted the usual scale of spaces associated to the operator H0 , that is H−n :=
hH0 in/2 H, where hH0 i := (1 + H02 )1/2 .
    Clearly, in the case I H is self-adjoint on DomH0 . We will see that in the case II one can easily
define H as a self-adjoint operator, but its domain is no longer equal to DomH0 . In the case III, strictly
speaking, the formula (5.18) does not make sense. Nevertheless, it is possible to define a renormalized
Aronszajn-Donoghue Hamiltonian. To do this one needs to renormalize the parameter λ. This procedure
resembles the renormalization of the charge in quantum field theory.
    Consider first the case I. We can compute its resolvent. In fact, for z 6∈ spH0 we define an analytic
function
                                      g(z) := −λ−1 + (h|(z − H0 )−1 h).                             (5.20)
Then for z ∈ Θ := {z ∈ C\spH0 : g(z) 6= 0} and λ 6= 0, the resolvent of the operator H is given by
Krein’s formula
                     R(z) = (z − H0 )−1 − g(z)−1 (z − H0 )−1 |h)(h|(z − H0 )−1 .             (5.21)
For λ = 0, we set Θ = C\spH0 and clearly
   The following theorem will describe how to define the Aronszajn-Donoghue Hamiltonian in all the
cases I, II and III:
Theorem 5.1 Assume that:
(A) h ∈ H−1 , λ ∈ R ∪ {∞} and let R(z) be given by (5.22) or (5.21) with g(z) given by (5.20),
or
(B) h ∈ H−2 , γ ∈ R and let R(z) be given by (5.21) with g(z) given by
                                                      32
    Another way to define H for the case h ∈ H−2 is the cut-off method. For all k ∈ N we define hk as
in (5.35) and fix the running coupling constant by
                                        −λ−1                     2 −1
                                          k := γ + (hk |H0 (1 + H0 )  hk )
and set the cut-off Hamiltonian to be
                                             Hk := H0 + λk |hk )(hk |.                            (5.23)
Then the resolvent for Hk is given by
                        Rk (z) = (z − H0 )−1 + gk (z)−1 (z − H0 )−1 |hk )(hk |(z − H0 )−1 ,       (5.24)
where
                                       gk (z) := −λ−1               −1
                                                                         
                                                   k + hk |(z − H0 )   hk .                       (5.25)
                                                                                  1
Note that λk is chosen in such a way that the renormalization condition (gk (i) + gk (−i)) = γ. holds.
                                                                                  2
It is easy to see that if H0 is bounded from below, then limk→∞ λk = 0. Again, the cut-off Hamiltonian
converges to the renormalized Hamiltonian:
Theorem 5.2 Assume that h ∈ H−2 . Then lim Rk (z) = R(z).
                                                    k→∞
   Let us assume that h is cyclic. Then the support of the spectral measure of h wrt H0 is spH0 . If
g(β) = 0 and β 6∈ spH0 , then H has an eigenvalue at β and the corresponding eigenprojection equals
                         1{β} (H) = (h|(β − H0 )−2 h)−1 (β − H0 )−1 |h)(h|(β − H0 )−1 .
                                                          33
5.4    Friedrichs Hamiltonian
Let H0 be again a self-adjoint operator on the Hilbert space H. Let  ∈ R and h ∈ H. The following
operator on the Hilbert space C ⊕ H is often called the Friedrichs Hamiltonian:
                                               "            #
                                                    (h|
                                          G :=                .                              (5.26)
                                                  |h) H0
Recall that expression the operators (h| and |h) are defined by
                                        H 3 v 7→ (h|v := (h|v) ∈ C,
                                                                                                      (5.27)
                                        C 3 α 7→ |h)α := αh ∈ H.
   We would like to describe how to define the Friedrichs Hamiltonian if h is not necessarily a bounded
functional on H. It will turn out that it is natural to consider 3 types of h:
    Clearly, in the case I G is self-adjoint on C ⊕ DomH0 . We will see that in the case II one can easily
define G as a self-adjoint operator, but its domain is no longer equal to C⊕DomH0 . In the case III, strictly
speaking, the formula (5.26) does not make sense. Nevertheless, it is possible to define a renormalized
Friedrichs Hamiltonian. To do this one needs to renormalize the parameter . This procedure resembles
the renormalization of mass in quantum field theory. Let us first consider the case h ∈ H. As we said
earlier, the operator G with DomG = C ⊕ DomH0 is self-adjoint. It is well known that the resolvent of
G can be computed exactly. In fact, for z 6∈ spH0 define the analytic function
                                                                                                    (5.31)
                               = γ + h|( 2(z−Hi−z0 )(i−H0 )
                                                            −      i+z
                                                              2(z−H0 )(−i−H0 ) )h
                                                      34
 (4) Q(z)∗ = Q(z).
Therefore, by [Ka], there exists a unique densely defined self-adjoint operator G such that Q(z) = (z −
G)−1 . More precisely, for any z0 ∈ Ω, DomG = Ran Q(z0 ), and if ϕ ∈ Ran Q(z0 ) and Q(z0 )ψ = ϕ, then
Gϕ := −ψ + z0 Q(z0 )ψ,
Proof. Let z ∈ Ω. It is obvious that Q(z) is bounded and satisfies (4). We easily see that both in the
case (A) and (B) the function g(z) satisfies
Inserting (5.33) into (5.34) we get 0 = (z − H0 )−1 f and hence f = 0. Now (5.33) implies α = 0, so
KerQ(z) = {0}.
   Using (2) and (4) we get (Ran Q(z))⊥ = KerQ(z)∗ = KerQ(z) = {0}. Hence 3) holds. 2
   Let h ∈ H−2 and γ ∈ R. Let us impose a cut-off on h. For k ∈ N we define
where 1[−k,k] (H0 ) is the spectral projection for H0 associated with the interval [−k, k] ⊂ R. Note that
hk ∈ H and hence both (hk | and |hk ) are well defined bounded operators. Set
where
                                      gk (z) := k + (hk |(z − H0 )−1 hk ).                                 (5.37)
                                                                                               1
Note that k is chosen such a way that the following renormalization condition is satisfied:   2   (gk (i) + gk (−i)) =
γ. Let us also mention that if H0 is bounded from below, then limk→∞ k = ∞.
Theorem 5.4 Assume that h ∈ H−2 . Then lim Qk (z) = Q(z), where Q(z) is given by (5.30) and g(z)
                                                 k→∞
is given by (5.31).
                                                       35
Proof. The proof is obvious if we note that lim k(z−H0 )−1 h−(z−H0 )−1 hk k = 0 and lim gk (z) = g(z).
                                           k→∞                                     k→∞
2
   Thus the cut-off Friedrichs Hamiltonian is norm resolvent convergent to the renormalized Friedrichs
Hamiltonian.
   Let us assume that h is cyclic. Then the support of the spectral measure of h wrt H0 is spH0 . If
β = g(β) = 0 and β 6∈ spH0 , then G has an eigenvalue at β. The corresponding projection equals
                                                                    (h|(β − H0 )−1
                                                                                         
                                   −2    −1         1
         1β (G) = (1 + (h|(β − H0 ) |h))                                                    .
                                              (β − A)−1 |h) (β − H0 )−1 |h)(h|(β − H0 )−1
2
    When the conditions of Theorem 6.2 are satisfied, then we say that the operator A has a compact
resolvent.
Theorem 6.3 (1) Let A be normal. Then A has a compact resolvent iff spA = spd A.
(2) Let A be bounded from below and self-adjoint. Then A has a compact resolvent iff µn (A) → ∞.
Theorem 6.4 Let f, g ∈ L∞     d
                        loc (R ), lim|x|→∞ f (x) = ∞ and lim|x|→∞ g(x) = ∞. Then
                                            H := f (x) + g(D)
has a compact resolvent.
                                                       36
Proof. Clearly, the functions f, g are bounded from below. Fix  > 0. For r > 0,let
Then for z ∈ C\R the operator (min(g(D), r) − r)(z − f (x) − r)−1 is compact. Hence also the operator
is compact. Thus
                                spess (Hr ) = spess (f (x) + r) ⊂ [r + inf f, ∞[.
Therefore, there exists N such that for n > N
µn (Hr ) ≥ r − + inf f.
But Hr ≤ H. Hence
                                              µn (Hr ) ≤ µn (H).
Therefore, for n > N , we have µn (H) ≥ R −  + inf f . Thus µn (H) → ∞. 2
Proof. ⇒ We know that for any n dim 1B(λ, n1 ) (A) = ∞. Therefore, we can find an orthonormal system
v1 , v2 , . . . such that vn ∈ Ran 1B(λ, n1 ) (A). The sequence v1 , v2 , . . . satisfies (6.38).
     ⇐ Let  > 0. We have
Hence
                                         cn := k1B(λ,) (A)vn k → 1.
Let
                                                      1
                                            ṽn :=      1B(λ,) (A)vn .
                                                     cn
Then kṽn k = 1, w− limn→∞ ṽn = 0 and ṽn ∈ Ran 1B(λ,) (A). Hence Span{ṽ1 , ṽ2 , . . .} is infinite dimen-
sional. Thus, 1B(λ,) (A) is infinite dimensional. 2
Definition 6.6 A sequence of vectors vn satisfying the conditions of the above theorem will be called a
Weyl sequence for λ and the operator A.
Theorem 6.7 (Weyl) Let A, B ∈ B(V) be normal and let B − A be compact. Then spess A = spess B.
Proof. Assume that λ ∈ spess A. Then there exists a Weyl sequence v1 , v2 , . . . for λ and the operator A.
We have limn→∞ (B − A)vn = 0. Hence v1 , v2 , . . . is a Weyl sequence for λ and the operator B. 2
                                                         37
Theorem 6.8 Assume that A, B are normal operators such that for some z0 6∈ spA ∪ spB the operator
(z0 − A)−1 − (z0 − B)−1 is compact. Then
spess A = spess B.
Proof. By the Weyl theorem spess (z0 − A)−1 = spess (z0 − B)−1 . Then we use Theorem 6.1 to normal
operators A and B. 2
(z0 − f (D) − g(x))−1 − (z0 − f (D))−1 = (z0 − f (D) − g(x))−1 g(x)(z0 − f (D))−1
References
[Da] Davies, E. B.: One parameter semigroups, Academic Press 1980
[Ka] Kato, T.: Perturbation theory for linear operators, Springer 1966
[RS1] Reed, M., Simon, B.: Methods of Modern Mathematics, vol. 1, Academic Press 1972
[RS2] Reed, M., Simon, B.: Methods of Modern Mathematics, vol. 4, Academic Press 1978
38