Tifr 26
Tifr 26
By
F.F. Bonsall
By
F.F. Bonsall
Notes by
K.B. Vedak
Frank F. Bonsall
iii
Contents
v
Chapter 1
i) d(x, y) ≥ 0, x, yεE
ii) d(x, y) = 0 ⇐⇒ x = y
1
2 The contraction mapping theorem
lim d(xn , x) = 0
n→∞
xn+1 = T xn (n = 0, 1, 2, . . .),
then lim xn = u and
n→∞
Kn
d(xn , u) ≤ d(x1 , xo )
1−K
where K is a Lipschitz constant for T .
xn+1 = T xn (n = 0, 1, 2, . . .)
We have
d(xr+1 , xs+1 ) = d(T xr , T xs ) ≤ Kd(xr , xs ) (1)
and so
d(xr+1 , xr ) ≤ K r (x1 , xo ) (2)
Given p q, we have by (1) and (2),
T u = lim xn+1 = u.
n→∞
4 The contraction mapping theorem
Kn
d(u, xn ) ≤ d(u, x p ) + d(x p , xn ) ≤ d(u, x p ) + d(x1 , xo ) for n < p, by
1−K
(3). Letting p → ∞, we obtain
Kn
d(u, xn ) ≤ d(x1 , xo )
1−K
Hence 6
d(ψ1 , ψ2 ) ≤ tMd(φ1 , φ2 )
As tM < 1, this proves that T is a contraction mapping. By the contrac-
tion mapping theorem, there exists φεE with T φ = φ i.e., with
Z x
φ(x) = yo + f (t, φ(t))dt
xo
u = lim (T k )n T xo = lim T (T k )n xo
n→∞ n→∞
k n
= T lim (T ) xo (by the continuity of ) T
n→∞
= T u.
1 n n
|(T n f1 )(x) − (T n f2 )(x)| ≤ |λ| M d( f1 , f2 )(x − a)n , (a ≤ x ≤ b)
n!
8 The contraction mapping theorem
Then
1 n n
d(T n f1 , T n f2 ) ≤ |λ| M (b − a)n d( f1 , f2 )
n!
This proves that all T n and in particular T , are continuous and, for
1
n sufficiently large |λ|n M n (b − a)n < 1, so that T n is a contraction
n!
mapping for n large. Applying the theorem, we have a unique f εE with
T f = f which is the required unique solution of the equation (1).
10 Definition 1.5. Let (E, d) be a metric space and ε > 0. A finite sequence
xo , x1 , . . . , xn of points of E is called an ε - chain joining x0 and xn if
d(xi−1 , xi ) < ε (i = 1, 2, . . . , n)
X
n
dε (x, y) = inf d(xi−1 , xi )
i=1
i) d(x, y) ≤ dε (x, y)
X
n X
n
dε (T x, T y) ≤ d(T ci−1 , T xi ) ≤ K d(xi−1 , xi )
i=1 i=1
dε (T x, T y) ≤ Kdε (x, y)
But in view of the observations made in the beginning of this proof, (1)
implies that
lim d(T n xo , u) = 0
n→∞
i) p(x) ≥ 0 (x ∈ E)
13
14 Fixed point theorems in normed linear spaces
17 Two norms are equivalent if and only if they define the same topology.
[For the proof refer to Dunford and Schwartz ([14], p. 62) or Day [13,
p.9]].
S = {x : ||x|| = 1}
λξn + µηn = 0 (n = 1, 2, . . .)
But
1 1
(x − y) = (x − z) + (x − z′ ),
2 2
1 1
so that ||x − y|| ≤ ||x − z || + ||x − z′ ||
′
2 2
1 1 1 1
Hence ||x − y|| = || (x − z) + (x − z′ )|| = || (x − z)|| + || (x − z′ )||
2 2 2 2
As the norm is strictly convex,
λ(x − z) + µ(x − z′ ) = 0
Definition 2.3. The mapping P of Lemma 2.1 is called the metric pro-
jection onto K.
18 Fixed point theorems in normed linear spaces
Proof. Let E denote the normed space and let T K ⊂ A, a compact subset
of K. A is contained in a closed convex bounded subset of E.
T (B ∩ K) ⊂ T (K) ⊂ A ⊂ B
so T (B ∩ K) is contained in a compact subset of B, K and there is no
loss of generality in supposing that K is bounded. If Ao is a countable
dense subset of the compact metric space A, then the set of all rational
linear combinations of elements of Ao is a countable dense subset of the
closed linear subspace Eo spanned by Ao and A ⊂ E0 . Then T (K ∩
E0 ) ⊂ T (K) ⊂ A, a compact subset of E0 , and K ∩ E0 is closed and
convex. Hence without loss of generality we may assume that K is a
bounded closed convex subset of a separable normed space E with a
strictly convex norm (Theorem 2.1).
Fixed point theorems in normed linear spaces 19
1
Given a positive integer n, there exists a -net T x1 , . . . , T xm say in 23
n
T K, so that
1
min ||T x − T xk || < (x ∈ K) (1)
1≤k≤n n
Let En denote the linear hull of T x1 , . . . , T xm . Kn = K ∩ En is a
closed bounded subset of En and therefore compact (Lemma 2.3). Since
the norm is strictly convex, the metric projection Pn of E onto the con-
vex compact subset Kn exists. T n = Pn T is a continuous mapping of the
non-empty convex compact subset Kn into itself, and therefore by the
Brouwer fixed point theorem, it has a fixed point un εKn ,
T n un = un (2)
|x − xo | ≤ ε, |y − yo | ≤ mε
| f (x, y)| ≤ m.
Remark . Theorem 2.2 and 2.3 are almost equivalent, in the sense that
Theorem 2.2, with the additional hypothesis that K be complete, follows
from Theorem 2.3. For, if K is a complete convex set and T K is con-
tained in a compact subset A of K, then the closed convex hull of A is a
compact convex subset Ko of K, and T K0 ⊂ K0 .
(ii) ||F(x, y) − F(x′ , y)|| ≤ ||Ax − Ax′ || (x, x′ , y ∈ K). Then there exists
a point u in K with
F(u, u) = u.
T x = F(x, T x) (x ∈ K).
We have ||T x − T x || = ||F(x, T x) − F(x′ , T x′ )||
′
1
Therefore ||T x− T x′ || ≤ ||Ax− Ax′ ||, (1) which shows that T k is
1−k
continuous and that T K ⊂ K is precompact since AK is compact, since
K is complete, T K ⊂ K is compact. By the Schander theorem, T has
fixed point u in K,
T u = u.
27 But then
F(u, u) = F(u, T u) = T u = u
αAu + (1 − α)Bu = u
28
p(xn − yn ) ≥ ε, (2)
and ! !
1 1
p (xn + yn ) > 1 − max(p(xn ), p(yn )). (3)
2 n
Let αn = p(xn ), βn = p(yn ), γn = max(αn , βn ). By (1) and (2),
1
γn ≥ , (4)
2
and so, by (1) and (3)
!
1 1
lim p (xn + yn ) = 1. (5)
n→∞ γn 2
Therefore, !
1 1
lim p xn − yn = 0.
k→∞ αnk k βnk k
and so
1
lim p(xnk − ynk ) = 0,
k→∞ γnk
In fact if y = T y , then
1 1
||xn+1 − y|| = || (xn + T xn ) − (y + T y)||
2 2
1 1
= || (xn − y) + (T xn − T y)||
2 2
1 1
≤ ||xn − y|| + ||T xn − T y||
2 2
≤ ||xn − y||
which is (1).
For
(||T x − x|| + ||x||)2 − ||T x||2 = ||T x − x||2 + ||x||2 − ||T x||2
+2||x|| ||T x − x|| ≥ 2r||T x − x|| > 0
Let T̃ = PT
Then T maps Q continuously into a compact subset of Q. Hence, by
the Schauder theorem, T̃ has a fixed point u in Q,
PT u = u
If T u ∈ Q, then PT u = T u, and
Tu = u
(1) For further results connected with Theorem 2.6, see [19]
x = λAx + y
has a solution x in E 33
Consider the mapping T x = λAx + y
28 Fixed point theorems in normed linear spaces
||T xn || ||T xn ||
> 1 so lim ≥1
||xn || ||xn ||→∞ ||xn ||
As this is not true, T maps some S n into its compact subset; Schau-
der’s theorem then a gives a fixed point x which is the required
solution.
P
∞
(3) Let ak be a convergent series of non-negative real numbers and
k=1
let( fk ) be
a sequence of continuous mappings of the real line R into
itself such that
| fk (t)| ≤ ak (t ∈ R, k = 1, 2, . . .)
(i) ξ = α
(ii) ξk+1 − ξk = fk (ξk ) (k = 1, 2, . . .)
(T x)1 = α
X
n
(T x)n+1 = α + fk (ξk ) (n = 1, 2, . . .)
k=1
34 where x = (ξk ).
Fixed point theorems in normed linear spaces 29
(4) Let E be a Banach space with a uniformly convex norm, and let K
be a bounded closed convex subset of E. Then the metric projection
onto
E −−−−→ K exists and is uniformly continuous on each bounded
subset of E.
(5) Brodsku and Milman [10], give conditions under which a convex
set in a Banach space has a point invariant under all isometric self
mappings. In this connection see also Dunford and Schwartz ([14],
p.459).
(x, y) → x + y
(α, x) → αx
31
32 The Schauder - Tychonoff theorem
(x, y) → αx + (1 − α)y
αa + (1 − α)b ∈ K0 + (1 − α)K0 ⊂ K,
v = h′ − h ∈ 2H ⊂ G. Thus V ⊂ G ⊂ U.
i) p(x) ≥ 0 (x ∈ E)
K = {x; pK (x) ≤ 1}
Given ε > 0,
x′ ε, x + εK ⇒ x′ − xεK
⇒ pK (x′ − x) ≤ ε
⇒ |pK (x′ ) − pK (x)| ≤ ε
p(x) ≤ 1 ⇔ x ∈ K ⇔ pk (x) ≤ 1,
V = {y : p(y) ≤ 1}
is a norm on E/N
Proof. If x, y ∈ N, then 40
H−H ⊂U
x′ ∈ x j + V(x j ) ⊂ x j + G(x j )
x − x j = x − x′ + x′ − x j ∈ V + V(X j ) ⊂ V(x j ) + V(x j ) ⊂ G(x j )
since x, x′ ∈ x j + G(x j ),
we have T x ∈ T x j + H, T x′ ∈ T x j + H,
and so T x − T x′ ∈ H − H ⊂ U.
We are now ready to prove the main theorem by which we are able to de-
duce properties of operators in a locally convex linear topological space
The Schauder - Tychonoff theorem 37
a) p′ (x) ≤ 1 (x ∈ K, p′ ∈ Γ′ ) 43
and so
XN
1 ǫ
q(x − x′ ) ⊂ n
pn (x − x′ ) + (x, x′ ∈ K) (2)
n=0
2 2
since T maps K into itself, (2) gives
XN
1 ǫ
′
q(T x − T x ) < p (T x − T x′ ) +
n n
(x, x′ ∈ K) (3)
n=0
2 2
since Γ is self-dominated, for each n, there exists kn and δn > 0 such that
ǫ
pkn (x − x′ ) < δn ⇒ pn (T x − T x′ ) < (x, x′ ∈ K) (4)
4
Let N ′ = max(k◦ , . . . , kN ), and
′
= 2−N min(δ◦ , . . . , δN ).
′
Then since pn ≤ 2N q for n ≤ N ′ , we have
px (T x − x) > 0
U x1 , . . . , U xm .
Let p = px1 + px2 + · · · + pxm .
E/N with the norm topology, K is a compact convex set in E/N. Also
by (2),
x, x′ ∈ K, q(x − x′ ) = 0 ⇒ q(T x − T x′ ) = 0 (3)
For each x̃ in K̃ there exists a point x in x̃ ∩ K, and we define T̃ x̃ by
taking
T̃ x̃ = Tfx
By (3), this definition is unambiguous, and T̃ maps K̃ into itself.
Also, by (2), given ε > 0, there exists δ > 0 such that x̃, x̃′ ∈ K, || x̃− x̃′ || <
δ ⇒ ||T̃ x̃ − T̃ x̃′ || < ǫ. For given x̃, x̃′ ǫK, there exist x ∈ K ∩ x̃ and 48
x′ ǫK ∩ x′ and q(x − x′ ) = || x̃ − x̃′ ||. Hence T̃ is a continuous mapping of
the compact convex subset K̃ of the normed space E/N. Applying the
Schauder fixed point theorem, T̃ has a fixed point ũ say
T̃ ũ = ũ.
Since ũǫ K̃, there exists u ∈ K ∩ ũ, and we have T̃ ũ = Tfu. Thus
T u − u ∈ N,
i.e., q(T u − u) = 0
TH ⊂ TK ⊂ H
The theorems in this chapter are mainly due to Krein and Rutman [20] 50
and to Schaefer [28]. They may be regarded as a further step in the
transition from nonlinear to linear problems. We will be content with
considering normed spaces only, through theorems of the kind studied
here have been proved for general locally convex spaces by H. Schaeffer.
(i) x, y ∈ C ⇒ x + y ∈ C
(ii) x ∈ C, α ≥ 0 ⇒ α ∈ C
(iii) x, −x ∈ C ⇒ x = 0
(v) x ≤ x (x ∈ E),
(vi) x ≤ y, y ≤ z ⇒ x ≤ z,
43
44 Nonlinear mappings in cones
(vii) x ≤ y, y ≤ x ⇒ x = y.
Also the partial ordering and the linear structure are related by the prop-
51 erties:
(viii) xi ≤ yi (i = 1, 2) ⇒ x1 + x2 ≤ y1 + y2 ,
(ix) x ≤ y, 0 ≤ α ≤ β ⇒ αx ≤ βy.
Conversely, given a non-trivial relation ≤ in E satisfying (v), . . . ,
(xi), the set {x : 0 ≤ x} is a positive cone in E to which the given relation
corresponds in the above manner.
1 1 1
x ∈ K, ||x|| ≥ c ⇒ ||Bx|| ≥ γc−1 cε = γε
2 2 2
On the other hand,
1 1 1
x ∈ K, ||x|| ≤ c ⇒ c−1 (c − ||x||)||y|| ≥ c ⇒ ||Bx|| γc
2 2 2
1
Therefore ||Bx|| ≥ γ min(ε, c) > 0 (x ∈ K).
2
It follows that the mapping
x → c||Bx||−1 Bx
46 Nonlinear mappings in cones
The following theorem due to Krein and Rutman [20, Theorem 9.1]
marks a further transition towards a linear problem.
54 (i) positive-homogeneous of
T (αx) = αT x (α ≥ 0, x ∈ C)
x, y ∈ C, x ≤ y ⇒ T x ≤ T y
(iii) || f || ≤ 1
Proof. Let
p(x) = d(x, −C) = inf{||x + y|| : y ∈ C}
Then p is a sublinear functional on E, with the properties
Nonlinear mappings in cones 47
T u ≥ cu (1)
Let Γ denote the set of positive real numbers t with x ≥ tu. Since
β 1 β β
x = u + T x ≥ u, we have ∈ Γ. Also Γ is bounded above,
α α α α
for otherwise
1
x ≥ u (n = 1, 2, . . .),
n
48 Nonlinear mappings in cones
For x ∈ Kε , x ≥ ε || x || u gives
T x ≥ ε || x || T u
By (3), || T x ||≥ f (T x)
and by (2) f (T x) ≥ f (εc || x || u) = εc || x || f (u) (xǫK∈ )
i.e., || T x ||≥ εc || x || f (u) ≥ εe f (x) f (u)
Nonlinear mappings in cones 49
≥ δε f (u) (x ∈ Kε )
Vǫ (0) = 0,
Vε (x) =|| x || . || x + 2ε || x || u ||−1 (x + 2ε || x || u), x , 0
|| x + 2ǫ || x || u ||≥|| x || −2ε || x || || u ||
=|| x || (1 − 2ε) > 0 if || x ||, 0.
|| Vǫ x ||=|| x || (7)
Also
x ∈ C, || x ||= 1 Vε x ∈ Kε (8)
For f (Vε x) = ||x||||x + 2ε||x||u||−1 { f (x) + 2ε||x|| f (u)}
2ǫ
≤ f (u) ≥ ε f (u)
1 + 2ǫ
Let Aε be the mapping defined on Kǫ by 58
Aε = Vε LT
x
when Lx = x,0
||x||
Then by (6) and (8)
Aε Kε ⊂ Kε
By (6), Vε L is continuous in T Kε , and Aε continuously into a compact
subset of Kε . Applying the Schauder theorem, we see that there exists a
point xε in Kε such that
Aε xε = xε ,
( )
Tx
i.e., Vε = xε (9)
||T xε ||
50 Nonlinear mappings in cones
i.e. ||T xε ||−1 T xε + 2εU = || ||T xε ||−1 T xε + 2εu||xε This can be written in
the form
T xε = αε xε − βε u, (10)
where c < αε < (1 + 2ε)||T xε ||.
We now choose a sequence (εn ) such that lim εn = 0, and such
n→∞
that the sequences (T xεn ) and (αεn ) converges. Let v = lim T xεn and
n→∞
59 λ = lim αǫn . Then λ ≥ c, and since lim n → ∞βεn = 0, (10) gives
n→∞
1
lim xεn = v
n→∞ λ
By continuity and positive homogeneity of T ,
!
1 1
T v = T v = lim T xǫn = v
λ λ n→∞
61
51
52 Linear mapping in cones
lim an = a
n→∞
λxnk → u,
λbxnk → bu,
E will denote a normed and partially ordered vector space with norm
||x|| and positive cone C. We suppose that C is complete with respect to
||x||, and that
E =C −C
We do not suppose that E is complete with respect to ||x||. We denote by
B the intersection of C and the closed unit of E, i.e.,
B = {x : x ∈ C and ||x|| ≤ 1}
B0 = {αx − βy : x, y ∈ B, α ≥ 0, β ≥ 0, α + β = 1},
||x||c ≤ 1 (x ∈ B),
and therefore
||x||c ≤ ||x|| (x ∈ C)
54 Linear mapping in cones
Therefore
wk+1 − wk ∈ 2−k B0 ,
and so wk+1 − wk = αk xk − βk yk , with
αk ≥ 0, β ≥ 0, αk + βk = 1, xk , yk ∈ 2−k B
Let
X
n X
n
sn = αk xk , tn = βk yk .
k=1 k=1
and similarly
||t p − tq || < 2−q .
65 Since C is complete, there exist s, t in C such that
||x|| ≤ ||x||c (x ∈ E)
and the closeness of C with respect to ||x||, (in fact a larger norm gives a
stronger topology).
x = αy − βz
67 with α ≥ 0, β ≥ 0, α + β = 1, y, z ∈ B. Then
T x = αT y − βT z,
and so
For the converse and the reversed inequality it is enough to note that
1
That µ = lim {||T n ||c } n
is an obvious consequence of the fact that
n→∞
||T ||c = p(T ) for each partially bounded linear operator.
If λ > µ, the series
1 1 1
I + T + 3 T2 + · · ·
λ λ2 λ
Linear mapping in cones 57
68 converges with respect to the operator norm for bounded linear operators
in the Banach space (E, ||x||c ) to a bounded linear operator Rλ , and
(λI − T )Rλ = Rλ (λI − T ) = I.
Since C is closed with respect to ||x||c , and the partial sums of the
series are obviously positive operators, it follows that Rλ is a positive
operator.
lim p(Rλ ) = ∞
λ→µ+0
λR − λx ≥ x (λ > 0, x ∈ C) (1)
If we let λ tend to zero through values for which p(Rλ ) ≤ M, the left
hand side of (1) tends to zero, and, since C is closed, we obtain
−x ∈ C (x ∈ C).
58 Linear mapping in cones
But this implies that C = (0) which was excluded by our axioms on C.
Suppose now that µ > 0. Then we may choose λ, ν with
and with p(Rν ) ≤ M. Since ||R||c = p(Rν ), it follows that the series
S (λI − T ) = (λI − T )S = I
70 This gives
S x = λ−1 x + λ−1 T S x (x ∈ C),
from which it follows by induction that
S x ≥ λ−(n+1) T n x (x ∈ C, n = 0, 1, 2, . . .) (2)
1
since λ < µ and lim ||T n ||cn = µ, we have
n→∞
as > anr−1 + r
Linear mapping in cones 59
Hence we see that there exists a strictly increasing sequence (nk ) of pos-
itive integers for which
and
||λ−(nk +1) T nk || ≥ ||λ−nk T nk −1 w|| (6)
since
||T nk w|| ≤ p(T )||T nk −1 w||,
we also have
lim ||λ−nk T nk −1 w|| = ∞ (7)
k→∞
Then, by (2),
But, by (6),
λ−1 ||T nk w|| ≥ ||T nk −1 w||,
which obviously contradicts (10). This contradiction proves the lemma.
To deduce the corollary, it is enough to appeal to the continuity of
the resolvent operator on the resolvent set.
Theorem 5.1. Let E be a normed and partially ordered vector space with
norm topology τN , positive cone C complete with respect to the norm,
and let B = {x : x ∈ C, ||x|| ≤ 1}. Let T be a partially bounded positive
linear operator in E with partial spectral radius µ, and let τ be a linear
73 topology in E with respect to which C is closed and T is continuous.
(ii) A = B,
then there exists a non-zero vector u in C with T u = µu.
αn x = T yn + zn
We have
T ynk + znk → 0 (τ),
and (T ynk ) has a τ-cluster point, v say, in C. It follows that −v is a τ- 74
cluster point of (znk ), and, since C is τ-closed, −v ∈ C, v = 0. This now
implies that 0 is a τ-cluster point of (T ynk ) and therefore is a τN -cluster
point of (T ynk ), which contradicts (1).
Thus, by Lemma 5.3, we have
lim p(Rλ ) = ∞,
λ→µ+0
i.e., lim ||Rλ ||c = ∞.
λ→µ+0
and we may suppose that ||Rλn w|| , 0 (n = 1, 2, . . .). Let αn = ||Rλn w||−1
and un = αn Rλn w. Then un ∈ B, ||un || = 1, lim αn = 0, and µun − T un =
n→∞
(µ − λn )un + (λn I − T )un = (µ − λn )un + αn w.
Suppose that condition (ii) in the statement of the theorem is satis-
fied. Since B is τ-countably compact and un ∈ B, it follows from (2)
that
lim µun − T un = 0 (τ) (3)
n→∞
75
Also (un ) has a τ-cluster point u in C, and (3) shows that µu−T u = 0.
We have u , 0, for otherwise a is 0 τN -cluster point of (un ), which
contradicts ||un || = 1.
Finally suppose that the condition (i) is satisfied. Then by (2).
(µ − I − T )T un = T (µI − T )un = (µ − λn )T un + αn T w
Example 1. Let E = CR [0, 1] with the uniform norm, and let C be the
positive cone in E consisting of those functions f belonging to E that
are increasing, convex in [0, 1] and satisfy f (0) = 0.
(T f )(x) = f (kx) ( f ∈ E, 0 ≤ x ≤ 1)
p(T n ) = kn ,
Also,
T fn = fn−1 (n = 1, 2, . . .),
and for r > s,
Tf = f ◦φ ( f ∈ E)
µ = φ′+ (0)
φn = T n φ (n = 1, 2, . . .).
1
and therefore lim ||T n φ|| n = φ′+ (0),
n→∞
µ ≥ φ′+ (0).
p(T n ) ≤ φn (1),
and so µ ≤ φ′+ (0). This completes the proof that Theorem 5.2 is appli- 80
cable.
In this particular example, we can calculate an eigenvector g by an
iterative process. In fact, if we take gn defined by
φn (x)
gn (x) = (0 ≤ x ≤ 1, n = 1, 2, . . .)
φn (1)
[0, 1] × [0, 1]
f (x) ≥ 0 (0 ≤ x ≤ 1),
Then T satisfies the condition of Theorem 5.2 except that its spectral
radius is zero (and hence its partial spectral radius is zero).
82 Also, if f ∈ C and T f = 0, we have
Z 1
K(1, y) f (y)dy = 0,
0
and so f = 0.
Linear mapping in cones 67
Theorem 5.3. Let X be a normed and partially ordered vector space with
a closed positive cone K, and suppose that there exists a subset H of K
with the properties:
f (T x) = µ∗ f (x) (x ∈ X).
Proof. Let X ∗ be the dual space of X, and let C be the dual cone K ∗
consisting of all f in X ∗ that satisfy
f (x) ≥ 0 (x ∈ K)
−h ≤ x ≤ h,
68 Linear mapping in cones
and therefore −T h ≤ T x ≤ T h,
− f (T h) ≤ f (T x) ≤ f (T h),
| f (T x)| ≤ f (T h) ≤ ||T h|| ≤ p(T )||h|| ≤ p(T ).M,
(T ∗ f )(x) = f (T x) (x ∈ X),
p(T ∗ ) ≤ M p(T ).
84 similarly
p(T ∗n ) ≤ M p(T n ),
and therefore
µ∗ ≤ µ
where µ∗ denotes the partial spectral radius of T ∗ . We take τ to be the
weak topology in E. Plainly C is τ-closed, B is τ-compact and T ∗ is
τ-continuous. In order to apply Theorem 5.1 it only remains to prove
that τ is sequentially stronger than τN at 0 relative to B. To prove this,
suppose that fn ∈ B(n = 1, 2 . . .) and that 0 is a τ-cluster point of the
sequence ( fn ).
Since H is contained in a (norm) compact set, given ε > 0. there
exists h1 , . . . , hr in H such that for each h ∈ H there is some k(1 ≤ k ≤ r)
with
ε
||h − hk || < (1)
2
Since 0 is a weak ∗-cluster point of ( fn ), there exists an infinite set
Λ of positive integers such that
ε
| fn (hk )| < (k = 1, 2, . . . , r; n ∈ Λ) (2)
2
Therefore by (1) and (2) and the fact that || fn || ≤ 1,
f (x) ≥ γ||x||
Therefore
86
Since, similarly,
p(T ∗n ) ≥ γp(T n ),
we have µ∗ ≥ µ,
Theorem 5.4 (Krein and Rutman ). Let X be a normed and partially or-
dered over space with a closed normal positive cone K with non-empty
interior. Let T be a positive linear operator in X. Then
(i) T is a bounded linear operator in X.
(ii) There exists a positive continuous linear functional f such that
f (T x) = ρ f (x) (x ∈ X),
where ρ is the spectral radius of T .
||x|| ≤ 1 e ± x ∈ K ⇒ −e ≤ x ≤ e (1)
Thus conditions (i) and (ii) of Theorem 5.3 are satisfied with
H = (e).
2||T e|| = ||(T e + T x) + (T e − T x)|| ≥ γ||T e − T x|| ≥ {||T x|| − ||T e||} ,
2+γ
and so ||T x|| ≤ ||T e||,
γ
which proves that T is bounded, and also gives
2+γ
||T || ≤ ||T e|| (3)
γ
It follows from (3) that
2+γ
||T || ≤ ||e|| p(T ),
γ
Linear mapping in cones 71
Theorem 5.5 (Krein and Rutman theorem 6.1). Let X be a partially or-
dered Banach space with a positive cone K such that X is the closed
linear hull of K. Let T be a compact linear operator in X that maps K
into itself and has a positive spectral radius ρ. Then there exists a non-
zero vector u in K and a positive continuous linear functional f such
that
T u = ρu, T ∗ f = ρ f.
Proof. Let ε > 0, and suppose that the lemma is false. Then,for each x
in X there exists a positive integer N x such that
ε
n ≥ N x ⇒ ||T n x|| < ||T n ||
2
89
Also
ε ε
||x′ − x|| < ⇒ ||T n x′ − T n x|| < ||T n ||,
2 2
72 Linear mapping in cones
ε
and so ||T n x′ || < ε||T n || (n ≥ N x , ||x′ − x|| < )
2
Let S denote the closed unit ball in X.Then T S is compact and so has a
finite covering by open balls of radius 2ε and centers x1 , . . . xm say. Let
N max(N x1 , . . . , N xm ).
Then
and so
||T n+1 || ≤ ε||T n || (n ≥ N),
from which it follows that lim ||T n ||1/n ≤ ε.
n→∞
T u = ρu
f (x) ≥ 0 (x ∈ K)
and let
M = (T − λI)γ X.
Then N and M are closed subspaces of X and
X=N⊕M (1)
T (M) ⊂ M, T N ⊂ N.
I = P + Q, PQ = QP = 0, P2 = P, Q2 = Q (2)
PX = M, QX = N.
(P + T P)x = αx.
α − 1 ∈ σ(T ) (ρ).
Therefore
Since all points ζ of σ(T ) satisfy |ζ| ≤ ρ, and since ρ is as isolated point 95
of σ(T ), it follows that
We have
I + T = (1 + ρ)I + A
and An Q = 0 (n ≥ ν).
Hence
( ! ! )
n n n n−1 n n−ν+1 ν−1
(I + T ) Q = (1 + ρ) I + (1 + ρ) A + · · · + (1 + ρ) A Q
1 −1
It follows that
!−1
−n n
lim (1 + ρ) (I + T )n Q = (1 + ρ)1−ν Aν−1 Q (2)
n→∞ ν−1
Also (1) gives,
!−1
−n n
lim (1 + ρ) (I + T )n P = 0 (3)
n→∞ ν−1
96 and, since (I + T )n = (I + T )n P + (I + T )n Q, (2) and (3) give
!−1
n
lim (1 + ρ)−n (I + T )n = (1 + ρ)1−ν Aν−1 Q (4)
n→∞ ν−1
Taking norms, we have
!−1
−n n
lim (1 + ρ) ||(I + T )n || = (1 + ρ)1−ν ||Aν−1 Q|| (5)
n→∞ ν−1
Linear mapping in cones 77
Qx = f (x)u,
i) (x, y) = (y, x)
79
80 Self-adjoint linear operator in a Hilbert space
Let H be a real or complex Hilbert space, and let S denote the class
of all bounded symmetric operators in H, i.e., bounded linear mappings
of T into itself such that
(T x, y) = (x, T y) (x, y ∈ H)
T ≥0
T A x = lim T n Ax = lim AT n x = AT x (x ∈ H)
n→∞ n→∞
99
Self-adjoint linear operator in a Hilbert space 81
T ⊂ E ′ ⇒AT = T A (A ∈ E)
∗ ∗
A T = TA (A ∈ E)
∗ ∗
T A = AT (A ∈ E)
(T ) ⊂ (T )′ ⇒ (T )′′ ⊂ (T )′
A1 , A2 ∈ (T )′ ⇒ A1 ∈ (T )′′ , A2 ∈ (T )′ A1 A2 = A2 A1 .
M = sup{(T x, x) : ||x|| ≤ 1}
and so ||T || ≤ M.
(T x, n) = 0 (x ∈ H)
By (b), this gives T = 0. The other properties of the cone are obvi-
ous.
T n ≤ T n+1 ≤ M.I (n = 1, 2, . . .)
lim T n x = T x (x ∈ H)
n→∞
(e) T ≥ 0 ⇒ T n ≥ 0 (n = 1, 2, . . .).
Self-adjoint linear operator in a Hilbert space 83
Proof. (T 2K x, x) = (T K x, T K x) ≥ 0
(T 2k+1 x, x) = (T.T x , T k x) ≥ 0.
1
(f) Each positive operator T is the square of a positive operator T 2 , and
1
T 2 belongs to the second commutant (T )′′ of T .
0 ≤ B ≤ I.
0 ≤ Yn ≤ I.
1 2 2 1
Yn+1 − Yn = (Yn − Yn−1 ) = (Yn + Yn−1 )(Yn − Yn−1 ),
2 2
we see by induction that Yn+1 −Yn is a polynomial in B with non-negative
real coefficients. Since Bn ≥ 0, for every n, it follows that (Yn ) is an in-
creasing sequence. Hence (Yn ) converges strongly to a positive operator
Y, and we have
1
0 ≤ Y ≤ I, Y = (B + Y 2 )
2
Let X = I − Y. Then X is a positive operator and
X 2 = A.
1
If 0 ≤ T ≤ MI, A = .T satisfies 0 ≤ A ≤ I and so the proposition
M
holds for T .
1
The positive square root T 2 is in fact unique but we do not need this
fact.
(g) A ≥ 0, B ≥ 0, AB = BA ⇒ AB ≥ 0.
1
Proof. Since A ∈ (B′ ), we have B 2 ∈ (A)′ and so
1 1 1 1
AB = AB 2 B 2 = B 2 AB 2 .
1 1 1 1
Finally, (B 2 AB 2 x, x) = (AB 2 x, B 2 x) ≥ 0.
(h) T ≥ 0 I + T is invertible, (I + T −1 ) ≥ 0, and
(I + T )−1 ∈ (T )′′ .
103
Proof. We have
I ≤ I + T ≤ (1 + M)I,
1
≤ A ≤ I,
1+M
1
where A = (I + T ). Therefore
1+M
!
1 M
||I − A|| ≤ || 1 − I|| = < 1.
1+M 1+M
I + (I − A) + (I − A)2 + · · ·
Self-adjoint linear operator in a Hilbert space 85
I = B(I − A) = I + (I − A)B = B,
and so AB = BA = I.
Finally (1 + M)−1 B is the required positive inverse of I + T . By a
projection we mean an operator P belonging to S with P2 = P.
P1 ≥ P2 ⇐⇒ P2 = P2 P1 ⇐⇒ P2 = P1 P2
104
and therefore P1 ≥ P2 .
Finally suppose that P1 ≥ P2 . If P1 x = 0, then P2 x = 0, for
Since P1 (I − P1 ) = 0,
P2 (I − P1 ) = 0,
i.e., P2 = P2 P1 .
86 Self-adjoint linear operator in a Hilbert space
A − B = (I + A2 )−1 (I + A2 )(A − B) ≥ 0
P = P2 ≤ PA ≤ A2 ,
and therefore
P = P2 ≤ A2 P.
Therefore
B − P ≥ (I + A2 )−1 2A2 (I − P) ≥ 0.
Theorem 6.1. Let A be a positive operator, and let the sequence (An ) be
defined inductively by
106 Then
Self-adjoint linear operator in a Hilbert space 87
i) 0 ≤ An+1 ≤ An (n = 1, 2, . . .),
iii) Q ≤ A,
iv) (I − A)(I − Q) ≥ 0,
Proof. (i) This follows at once from Lemma 1. (ii) and (iii). It follows
from (i) and Proposition (d) that (An ) converges strongly to a positive
operator Q with Q ≤ A, and that Q ∈ (A)′′ . It remains to prove that Q is
a projection.
Since 0 ≤ An ≤ A, we have
lim An x = Qx (x ∈ H),
n→∞
we have in turn,
But 107
Therefore
(Q − Q2 )2 = 0.
88 Self-adjoint linear operator in a Hilbert space
(Q − Q)2 = 0,
i.e., Q is a projection.
(iv) By Lemma 1 (iii),
mI ≤ T ≤ MI.
iii) Eλ ≤ Eµ (λ ≤ µ);
1
Eλ ≤ (MI − T ),
M−λ
and so
1
(x0 , x0 ) = (Eλ x0 , x0 ) ≤ ((MI − T )x0 , x0 ).
M−λ
Since MI − T ≤ (M − m)I, this gives
M−m
(x0 , x0 ) ≤ (x0 , x0 ),
M−λ
and so λ ≥ m. This proves (ii). 109
(iii) This is obvious except when λ < µ < M. In this case, since
1
Eλ is a projection permutable with (MI − T ), and
M−µ
1 1
Eλ ≤ (MI − T ) ≤ (MI − T ),
M−λ M−µ
Then 1 (v) shows that
Eλ ≤ Eµ
(T − λI)(Eµ − Eλ ) ≥ 0,
90 Self-adjoint linear operator in a Hilbert space
which is the left hand inequality in (iv). The right hand inequality is
obvious if µ = M ( since T ≤ MI); and if µ < M we have
1
Eµ ≤ (MI − T ),
M−µ
i.e., T ≤ MI − (M − µ)Eµ .
Since Eµ (Eµ − Eλ ) = Eµ − Eλ , this gives.
T (Eµ − Eλ ) ≤ µ(Eµ − Eλ ),
110 and (iv) is proved.
(v) Suppose µ < M. If (Eλ ) is not strongly continuous on the right at
µ, there exists a sequence (λn ) convergent decreasing to µ but such
that Eλn does not converge strongly to Eµ .
Since (Eλn ) is a decreasing sequence of operators, with Eλn ≥ Eµ ,
there exists a positive operator J such that J ≥ Eµ and (Eλn ) converges
strongly to J. Then J is a projection, permutable with T , and
1
Eµ ≤ J ≤ Eλn ≤ (MI − T ) (n = 1, 2, . . .)
M − λn
It follows that
1
Eµ ≤ J ≤ (MI − T ),
M−µ
and so by the maximal property of Eµ ,
J ≤ Eµ .
This completes the proof of the theorem.
Corollary. (The spectral theorem).
ZM
T = λdEλ (ε > 0)
M−ε
Then
Hence
X
n
||T − λk−1 (Eλk − Eλk−1 )|| → 0 as max(λk − λk−1 ) → 0
k=1
Moreover
ZM
r
T = λr dEλ (r = 0, 1, 2, . . .)
m−ε
To see this we rewrite (iv) in the form
ZM
r
(MI − T ) = (M − λ)r dEλ (r = 0, 1, 2, . . .)
m−ε
Lemma 6.2. Let A be a positive operator, and let (An ) be the sequence
constructed as in Theorem 7. Then An = A2 Bn (n = 2, 3, . . .), where
each Bn belongs to (A)′′ and
0 ≤ Bn+1 ≤ Bn (n = 2, 3, . . .)
Proof. We have
A2 = 2A2 (I + A2 )−1 = A2 B2 ,
with B2 = 2(I + A2 )−1 . If An = A2 Bn with Bn ≥ 0 and Bn ∈ (A)′′ , then
Theorem 6.3. Let A be a compact positive operator, and let (An ) and Q
be the corresponding and projection defined as in Theorem 1. Then
Q = A2 B.
Self-adjoint linear operator in a Hilbert space 93
Let K be the unit ball in the Hilbert space H, and let E = (AK).
Then E is a norm compact set. Since lim Bn x = Bx (x ∈ H), we have
n→∞
Px = (x, u)u.
APx = (x, u)Au = λ(x, u)u = (x, λu)u = (x, Au)u = (Ax, u) = PAx. Also
P ≤ A, for given x ∈ H, we have x = H, we have x = ξu + v with
(u, v) = 0. Then Ax = λξu + Av, and
(u, Av) = (Au, v) = λ(u, v) = 0. So
94 Self-adjoint linear operator in a Hilbert space
In this brief chapter we are concerned with the existence of a simultane- 116
ous fixed point of a family F of mappings
T u = u (T ∈ F ).
We first state without proof two well known theorems on this question
proofs of which will be found in Dunford and Schwartz [14] pp.456-
457. We then prove a theorem on families of mappings of a ∞ne into
itself. Some further results in the present context will appear in the next
chapter in the theory of a special class of semialgebras.
95
96 Simultaneous fixed points
(c) If E has an order unit e, and J is a proper ideal of E, then E/J is a 119
partially ordered vector space with order unit.
x + j, − x + j′ ∈ C.
This gives
0 ≤ x + j ≤ j′ + j
and so x + j ∈ [0, j + j] ⊂ J,
x ∈ J, x̃ = 0.
98 Simultaneous fixed points
−λe ≤ x ≤ λe
implies that − λẽ ≤ x̃ ≤ λẽ
120 (d) If E has an order unit, and M is a maximal proper ideal of E, then
E/M has no proper ideals.
Let
p(x) = inf[ξ : x ≤ ξe] (x ∈ E) (1)
Let y = p(x)e − x.
If x ∈ E, then either x ∈ C or −x ∈ C, for otherwise (x) is an ideal.
Hence y ∈ C or −y ∈ C i.e. y ≥ 0 or y < 0. If y < 0 for some x, then
−y ≥ εe for some ε > 0, and x ≥ (p(x) + ε)e which contradicts (1). If
y > 0, for some x ∈ E, then y is an order unit and so
y = p(x)e − x = 0 x ∈ E)
(f) Let E have an order unit e, and let M be a maximal proper ideal.
Then there exists a linear functional f on E with
(g) Let E have an order unit and have dimension greater than one, and
let T be a positive linear mapping of E into itself.
Then there exists a proper T -invariant ideal, i.e. a proper ideal J
with T J ⊂ J.
Hence, if 0 ≤ x ≤ j, then
−εe ≤ x ≤ εe (ε > 0)
−εe ≤ x ≤ εe (ε > 0)
100 Simultaneous fixed points
and so − εT e ≤ T x ≤ εT e (ε > 0)
But
Hence
αp(T x) = 0 i.e. T x ∈ N.
−p(x)e ≤ x ≤ p(x)e (x ∈ E)
so that − p(x) f (e) ≤ f (x) ≤ p(x) f (e) (x ∈ E, f ∈ H)
i.e., − p(x) ≤ f (x) ≤ p(x) (x ∈ E, f ∈ H)
or | f (x)| ≤ p(x)
1
Sf = T∗ f
f (T e)
Simultaneous fixed points 101
(h) Under the condition of (g) there exists a maximal proper ideal M
and a non-negative real number such that
T x − µx ∈ M (x ∈ E).
T̃ x̃ = T̃ x,
x − f (x)e ∈ M (x ∈ E),
and so T x− = f (x)T e ∈ M (x ∈ E)
f {T x − f (x)T e} = 0 (x ∈ E)
f (T x − µx) = 0 (x ∈ E),
and so T x − µx ∈ M (x ∈ E).
A class of abstract
semi-algebras
103
104 A class of abstract semi-algebras
(iii)′ A with the relative topology induced from the norm topology in
B is a locally compact space.
This is our justification for the use of the term locally compact in the
present sense. Axiom (ii) of course merely excludes trivial exceptional
cases.
If the Banach algebra B has finite dimensions, then its closed unit
ball is compact, and therefore every nontrivial closed semi-algebra in
B is locally compact. In particular, each closed semi-algebra of n × n
real matrices is of this kind. However, the axioms do not imply that
every locally compact semi-algebra is contained in a finite dimensional
algebra, as the following example shows.
Example. Let E be the subset of the closed unit interval [0, 1] consisting
1
of the closed interval [0, ] together with the point 1, and let E be given
2
the topology induced form the usual topology in [0, 1], so that E is a
compact Hausdorff space.
Let A′ denote the class of all functions belonging to CR [0, 1] that are
non-negative, increasing, and convex in [0, 1]; and let A denote the class
of all functions on E that are restrictions to E of functions belonging to
A′ .
It is obvious that A′ is a semi-algebra in CR [0, 1]. We prove that A
128 is a closed subset of CR (E). Each element f of A has a unique extension
1
f ′ ∈ A′ which is linear in [ , 1] defined by
2
1
f ′ (x) = f (x), 0 ≤ x ≤
! 2
1 1
f ′ (x) = α f + (1 − α) f (1) for x = α. + (1 − α)1 , 0 ≤ α ≤ 1
2 2
A class of abstract semi-algebras 105
Thus the set of all such f is equi-continuous, and A is a locally compact 129
semi-algebra. Finally it is obvious that A is not contained in any finite
dimensional subspace of CR (E).
Our principal results are concerned with the existence and properties
of idempotents in a locally compact semi-algebra, and may be regarded
as analogues of classical theorems of Wedderburn. As biproducts we
obtain an abstract characterization of the semi-algebra of all n × n ma-
trices with non-negative entries, and some results related to the Perron-
Frobenius theorems.
Throughout this chapter we will denote by S A the intersection of A
with the surface of the unit ball in B, i.e.,
Er = {x : x ∈ A and ux = 0 (u ∈ E)}.
It is clear that similar results hold for left and two sided closed ide- 131
als.
Hence
|| asni || > m > 0 (i = 1, 2, . . .).
Also since
lim ||xni ||asni = lim axni = y,
i→∞ i→∞
||xn ||asni is a bounded sequence. It follows that the sequence (||xni ||) is
bounded, and therefore has a subsequence convergent to λ > 0 say. Then
y = λ as = a(λs) ∈ aE
Theorem 8.2. Let M be a minimal closed right ideal of a locally com- 132
pact semi-algebra A with M 2 , (0). Then M contains an idempotent e
and M = eA.
108 A class of abstract semi-algebras
M ∩ (a)r , M
0 , aM ⊂ M
and therefore
aM = M.
In particular, there exists an element e ∈ M with
ae = a (2)
aen = a (n = 1, 2, . . .)
||a||||e || ≥ ||aen || = ||a||,
n
||en || ≤ 1 (n = 1, 2, . . .)
so that
lim ||en ||1/n ≥ 1
n→∞
In order to prove that lim ||en ||1/n ≤ 1, it suffices to show that (||en ||) is
n→∞
bounded.
Let K = inf{||am|| : m ∈ M ∩ S A }.
A class of abstract semi-algebras 109
1 1
bλ = e + 2 e2 + . . .
λ λ
The convergence of the series is established by (3), and we have
bλ ∈ M. Also,
λbλ − ebλ = e ∈ M,
and therefore bλ , 0. Let λn > 1(n = 1, 2, . . .) and lim λn = 1. By what 134
n→∞
we have just proved, there exists for each n, an element mn of M ∩ S A
such that
λn mn − emn ∈ M
Therefore, by the compactness of M ∩ S A , there exists an element m of
M ∩ S A such that
m − em ∈ M
Let
J = {x : x ∈ M, x − ex ∈ M}.
We have J , (0) since m ∈ J. Also, J is a closed right ideal con-
tained in M, and therefore J = M i.e.,
x − ex ∈ M (x ∈ M)
But, by (2)
a(x − ex) = 0 (x ∈ M),
and so
x − ex ∈ M ∩ (a)r (x ∈ M)
110 A class of abstract semi-algebras
Therefore, by (1),
x − ex = 0 (x ∈ M)
M = eM = eA.
A0 = {x : x ∈ A and x = ex = ex}.
(eae)r ∩ eA = (0).
x = x(yz) = (xy)z,
A = R+ e,
where e is the unit element of A and R+ is the set of all non-negative real
numbers.
e + z + z2 + · · ·
e − z = (e − z)ab = b ∈ A.
137
Finally
y − x = y(e − y−1 x) = y(e − z) ∈ A.
Suppose now that u ∈ A, u , 0, and let
µ = sup{λ : e − λu ∈ A}.
112 A class of abstract semi-algebras
(e − µu) − λu ∈ A,
i.e., e − (µ + λ)u ∈ A,
A = R+ e.
f (t) = 0 (t ∈ E),
or f (t) > 0 (t ∈ E).
Proof. We first show that if J is any left ideal with J 2 = (0) then
J = (0) (1)
A21 ⊂ A1 A = (0),
we haveA1 = (0), J = (0)
If now I is a left ideal and n is the least positive integer with I n = (0),
then I n−1 I = 0, and so n > 1 would give (I n−1 )2 = (0), and so by (1)
I n−1 = 0
Hence n = 1, I = (0).
A similar argument applies to right ideals.
eA = {x : x ∈ A and x = ex},
( f A)2 ⊂ ( f A)(eA),
f Ae , 0
f aeb = e.
Suppose ae , 0. Then e < (ea)r , and therefore the closed right ideal
(ea)r ∩ eA
(ae)r ∩ eA = (0)
aeA = (ae)(eA)
A class of abstract semi-algebras 115
ae = f b for some b ∈ A.
Proof. By Theorem 8.1 and the fact that A is a non-zero closed two-
sided ideal of itself, A has at least one minimal closed two-sided ideal.
Mα ∩ Mβ = (0) (α , β),
and so Mα Mβ (0) (α , β).
142
For each α ∈ ∆, choose mα ∈ Mα ∩ S A . By the compactness of S A ,
there exists a sequence (αn ), of distinct elements of ∆ such that (mαn )
converges to an element m say of S A . Given α ∈ ∆, we have
Ik A = AIk = Ik AIk ,
eAMk ⊂ eA ∩ Mk ,
S
n
we have eAMk = (0), e ∈ (Mk )l . Thus if e ∈ I but e < Ik , then
k=1
\
n
e∈ (Mk )l = J.
k=1
Mk ∩ (Mk )l , (0),
(u)r ∩ f A = (0)
eA = u f A ⊂ J,
J = eA
118 A class of abstract semi-algebras
eA f = R+ ω,
and either ω2 = ω or ω2 = 0.
veA = f A,
veA f = f A f.
ex f = ex f vu = λeu = λu
f A ⊂ Mj
and so, if Mi , M j ,
f Ae ⊂ M j Mi = (0)
eA f = R+ ωe, f ,
120 A class of abstract semi-algebras
Proof. Minimal closed two-sided ideals annihilate each other, and there-
fore if A is prime there is exactly one such ideal.
Suppose on the other hand that A is semi-simple and has exactly one
minimal closed two-sided ideal, and let H, J be closed two-sided ideals
with H J = (0). Then (H ∩ J)2 = (0) and so, by semi-simplicity of A,
H ∩ J = (0)
Proof. Since A has a unit element, A2 , (0). But A is the only non-zero
two-sided ideal, and therefore A is semi-simple and indeed prime. Let
I denote the class of all minimal idempotents, and let 1 be the unit
element. Then I A is a non-zero, two-sided ideal in A, and so
A = I A.
e1 = e1 a1 e1 + e2 a2 e1 + · · · + en an e1 (2)
1 = e2 b2 + · · · + en bn ,
(λ − 1)e1 + e2 a2 e1 + · · · + en an e1 = 0
λ = 1, e j a j e1 = 0( j , 1).
ui Au j = R+ ei j
eii = ui (i = 1, 2, . . . , n) (6)
ei j e jk = eik (i, j, k = 1, . . . , n), (7)
ei j ekl = 0 ( j , k) (8)
151
In the first place we have ui Aui = R+ ui , and so we can take eii =
ui (i = 1, 2, . . . , n). Next, for j = 2, . . . , n we take ei j to be an arbitrary
non-zero element of u1 Au j . Then we have
u1 Au j = R+ ei j ( j = 1, . . . , n).
e1 j e jk = e1k ( j, k = 1, . . . , n) (9)
A class of abstract semi-algebras 123
ei j e jk ∈ ui Auk
and so ei j e jk = λeik with λ ≥ 0.
X
n
y= ηi j ei j
i, j
n
X Xn
and xy =
ξi j η jl
eil .
i,l=1 j=1
so that
m
urs aulm = α sl ur .
It follows that every non-zero two-sided ideal of A contains all the
matrices urs , and so is the whole of A. Thus A is simple and the proof is
complete.
In the case when A is commutative our theorems on idempotents
take a particularly simple form. In the first place we can determine semi-
simple and prime commutative semi-algebras by annihilation properties
of individual elements.
a ∈ A, a , 0 ⇒ an , 0 (n = 2, 3, . . .)
154 Conversely if
a ∈ A, a2 = 0 ⇒ a = 0,
Then A is semi-simple.
K J = 0 ⇒ J ∩ (a)r , (0)
For some J, we have K J > 0. For otherwise every non-zero closed ideal
J satisfies J ∩ (a)r , (0), and therefore
M ∩ (a)r = M.
for every minimal closed ideal M i.e., aM = (0) for every such M. Let 155
I = cl(aA). Then I is a non-zero closed ideal and I M = (Q). for every
minimal closed ideal M. Since I contains a minimal closed ideal, this
would contradict the semi-simplicity of A.
Let J be a closed ideal with K J > 0. Then
Therefore
aek = λa ek (a ∈ A)
156 Proof. All ideals of A are two-sided, and so, by Theorem 8.7 A has only
finitely many minimal closed ideals M1 , . . . , Mn and we have
Mi M j = (0) i , j.
aei = λi ei (i = 1, 2, . . . , n),
and max {λi : 1 ≤ i ≤ n} = ρa
where ρa = lim ||ak ||1/k , and ρa > 0 (a , 0)
k→∞
Corollary. If also A is prime then there exists exactly one minimal idem-
potent e and
ae = ρa e (a ∈ A).
A class of abstract semi-algebras 127
T
n
Also (ei )l is a closed ideal, and so if it were non-zero it would
i=1
contain one of the minimal idempotents ei , which is absurd. Therefore
(e)l = (0).
Let a ∈ A,
K = inf{||xe|| : x ∈ S A }, λ = max(λ1 , . . . , λn ),
µ = ||e1 || + . . . + ||en ||
||x|| ≤ K −1 ||xe|| (x ∈ A)
ak e = λk1 e1 + · · · + λkn en ,
and so ||ak e|| ≤ λk µ.
Therefore
||ak || ≤ K −1 λk (k = 1, 2, . . .)
On the other hand, for some i we have λ = λi , 158
ak ei = λk ei ,
and therefore ||ak || ≥ λk
x = (ξ1 , . . . , ξn ) = ξ1 u1 + · · · + ξn un ,
aC∆ ⊂ C∆ .
i < ∆, j ∈ ∆ αi j = 0,
for it is equivalent to
au j ∈ C ( j ∈ ∆),
and au j is the vector (αi j , α2 j , . . . , αn j ). For example, if ∆ = (1, 2, . . . , r),
then (αi j ) is of the form !
b c
o d
where b is an r × r matrix.
A matrix a ∈ Mn (R+ ) is said to be irreducible if it is not reducible.
Given a subset E of Mn (R+ ), let
I J = (0)
a0 ∈ A ⊂ Mn (R+ ).
X = (u) + Y
130 A class of abstract semi-algebras
A = Aχ
A = Aχ = R+ χ,
E( f, α) = {x : f (x) ≥ α},
We have α > 0, and since χ f,α ∈ A, E( f, α) is both open and closed. But
t ∈ E( f, α) and s < E( f, α), and so E0 is not connected.
Let △ denote the set of all characteristic functions that belongs to
A. Then △ is a finite set, for if (χn ) were an infinite sequence of distinct
elements of △, it would have a uniformly convergent subsequence which
is absurd, since
||χ p − χq || = 1 (p , q)
We show that each element of A is a non- negative linear combination
of the elements of
A class of abstract semi-algebras 133
hi = χ f,αi (i = 1, . . . , n),
In Chapter 3, I asked whether the Schauder fixed point in its full gen- 167
erality (Theorem 2.2) is true for locally convex spaces, and pointed out
that this question did not seem to be answered in the literature. I am very
much indebted to B.V. Singbal who showed that this question could be
settled affirmatively by using a technique due to Nagumo. The result-
ing proof of the general Schauder fixed point theorem is in my view the
simplest proof even for the special case of normed spaces.
S x − x ∈ V (x ∈ A).
135
136 A class of abstract semi-algebras
PW (x − ai ) < 1.
For each x in A, there is at least one i with qi (x) > 0. Since also
qi (x) ≥ 0 for all i, it follows that S is defined and continuous on A
and that it maps A into the convex hull of a1 , . . . , am . For any i with
pw (ai − x) ≥ 1, we have qi (x) = 0, and therefore
pW (S x − x) < 1 (x ∈ A),
i.e., S x − x ∈ W ⊂ V (x ∈ A).
S V x − x ∈ V (x ∈ A). (1)
A class of abstract semi-algebras 137
139
140 BIBLIOGRAPHY
[31] J. Schauder, Zur Theorie stetiger Abbildun gen in Function alrau- 174
men, Math. Z. 26 (1927), 47-65, 417-431.