Quantizing Analysis
Quantizing Analysis
Kelvin Lui
12/11/2014
Department of Mathematics
Colby College
2015
Definitions
Definition 1. A C -algebra, A, is a Banach algebra over the field of complex numbers with a map , such that
x = (x ) = x
(x + y) = x + y
(xy) = y x
(x) = x
kx xk = kxk2
(1.0.1a)
(1.0.1b)
(1.0.1c)
(1.0.1d)
(1.0.1e)
(1.0.2a)
(1.0.2b)
(1.0.2c)
2
2.1
Abstract Algebra
Rings and Ideals
an tn .
(2.1.2)
n=0
an t +
an tn
M
X
n=0
M
X
n=0
max{N,M }
n
bn t =
bn tn =
(an + bn )tn
n=0
NX
+M X
n
n=0
ak bnk tn
(2.1.3a)
(2.1.3b)
k=0
and with this we see that R[t] is a ring. The set of all sequences (an )nN forms a
ring as well with the same definitions of addition and multiplication. It is called the
ring of formal power
P seriesn in t and denoted as R[[t]]. We view elements of R[[t]] as
infinite sums , n+0 an t .
Definition 11. If R is a ring, then an element a R\{0} is a zero-divisor if there
is some b R\{0} such that a.b = 0.
Definition 12. Let f : R S be a ring homomorphism. The kernel of f is
ker(F ) = {r R : f (r) = 0},
(2.1.4)
(2.1.5)
The image of a homomorphism is a subring of the target ring. The kernel, however
is not a subring of the target ring. The kernel is closed under addition and multiplication and since 0.x = 0, x, it obeys a stronger kind of closure with respect to
multiplication. If x ker(f ) and r R then f (x.r) = f (x).f (r) = 0.f (r) = 0 so
that x.r ker(f ).
Definition 13. Let R be a ring. A subset I R is called an ideal if it is a subgroup
of (R, +) and moreover for any a I and r R then a.r I.
Lemma 2.2. If f : R S is a homomorphism, then ker(f ) is an ideal.
I + J = {i + j : i I, j J};
IJ = {
n
X
ik jk : ik I, jk J, n N }.
(2.1.6)
k=1
ik jk +
k=1
m
X
max(n,m)
pk qk =
k=1
ik jk + pk qk
k=1
ik , pk I, jk , qk J, n, m N,
P
which is of the same form. If nk=1 xk yk IJ then
n
X
k=1
x k yk =
n
X
(2.2.4)
(2.1.8)
k=1
J =0+J I +J
(2.1.9)
2.2
Quotients
We will show later that ideals and kernels of ring homomorphisms are the same,
however we first need to know the notion of quotient for rings.
Let us suppose that R is a ring and I is an ideal in R. Since (I, +) is a subgroup of
the abelian group (R, +), we may form the quotient group (R/I, +); the construction
of the group being: r s if r s I, and the equivalence classes are the cosets
{r + I : r R}. To endow R/I with the structure of an abelian group suppose that
C1 , C2 R/I then we define
C1 + C2 = {c1 + c2 : c1 Ci , i = 1, 2}
(2.2.1)
(2.2.2)
(2.2.3)
We also claim that the set C1 C2 is contained in a single coset of R/I, that is
C1 C2 + I is a single I-coset. Suppose the ci , di are given as above, then
c1 c2 = (d1 + k1 ) (d2 + k2 ) = d1 d2 + (d1 k2 + k1 22 + k1 k2 ) d1 d2 + I (2.2.4)
{z
}
|
I
(2.2.5)
(2.2.6)
(2.2.7)
(2.2.8)
(2.2.9)
(2.2.10a)
(2.2.10b)
(2.2.11)
Proof. Let qI : R R/I and let qJ : R R/J be the two quotient maps. By the
universal property for the quotient qJ : R toR/J/ applied to the homomorphism qI we
see that there is a homomorphism qI : R/J R/I induced by the map qI : R R/I
and q qJ = qI . Clearly qI is surjective since qI , and if q(r + j) = 0 then since
r + J = qJ (r) so that qI (r + J) = qI (r) we have r I so that ker(
qI ) = J/I and the
result follows from the first homomorphism theorem.
2.3
Modules
(2.3.1a)
(2.3.1b)
(2.3.1c)
(2.3.1d)
(2.3.2)
(2.3.3a)
(2.3.3b)
(2.3.4)
The right hand side is simply writing the submodule S + N , S/N . However, if
N S then we have S + N = S and therefore q 1 (q(S)) = S and any submodule S
which contains N is indeed the preimage of a submodule M/N .
Theorem 2.11. Universal property of quotients
Suppose that : M N is a homomorphism of R-modules, and S is a submodule
of M with S ker(). Then there is a unique homomorphism : M/S N such
where q : M toM/S is the quotient homomorphism, that is the following
that = q
diagram commutes:
M
q
M/S
is the submodule ker()/S = {m + S : m ker()}.
Moreover, ker()
(m1 + S) + (m2 + S) = m1 + m2 + S
= (m1 + m2 )
= (m1 ) + (m2 )
1 + S) + (m
2 + S)
= (m
r.(m + S) = (r.m
+ S)
= (r.m) = r.(m)
+ S)
= r.(m
(2.3.5a)
(2.3.5b)
is m + S ker()/S.
As we have done above, we will now look at the isomorphism theorems for modules.
First we let M be an R-module.
Theorem 2.12. First Isomorphism theorem
If : M N is a homomorphism then induces an isomorphism : M/ker()
im().
Proof. Let K = ker(). Using the universal property of quotients to K, then we have
= ker()/ker() = 0 it follows that is injective and therefore induces an
ker()
isomorphism onto its image which from the equation q = = Im().
Theorem 2.13. Second isomorphism theorem
If M is an R-module and N1 , N2 are submodules of M then
(N1 + N2 )/N2
= N1 /N1 N2
(2.3.6)
Proof. First, let q : M M/N2 be the quotient map, which restricts p from N1 to
M/N2 , and whose image is clearly (N1 + N2 )/N2 . Therefore by the first isomorphism
theorem it is only necessary to check that the kernel of p is N1 N2 . If n N1 has
p(n) = 0 then n + N2 = 0 + n2 so that m N2 and so n N1 N2 .
Theorem 2.14. Third isomorphism theorem
Suppose that N1 N2 are submodules of M . Then we have
(M/N1 )/(N2 /N1 )
= M/N2
(2.3.7)
3
3.1
The study of dual spaces is essential to understanding the notion of a tensor product.
To formalize our understanding of the dual space we will look at real-valued functions
defined on a real vector space T , f : T R. The set of all such functions is then
given the structure
v T
v T
v T
v T,
(3.1.1a)
(3.1.1b)
(3.1.1c)
(3.1.1d)
which defines a vector space. Furthermore, we will restrict the space of all realvalued functions to only those that are linear
f (u + v) = f (u) + f (v) , R, u, v T.
(3.1.2)
The real-valued linear functions on a real vector space are also called linear functionals and the space constitutes a vector space called the dual of T , denoted by
T .
For simplicity we will point out the two types of vectors, those in T an those in
T , and distinguish them from one another. For the basis vectors of T we will use
superscripts and the components of vectors we will use subscripts, therefore
if {ea }
P
is a basis of T , then a vector T has the unique expression = a a ea .
In a natural way, we will now prove that the that the dual space of T has the
same dimensionality as T .
Lemma 3.1. The dual space,T , of T has the same dimensionality as T .
Proof. First, let T be a N -dimenionsal vector space and let {ea } be the given basis
of T . We define {ea } to be real-valued functions which maps T into the real
number a , which is the ath component relative to the basis vectors of T , {ea }. They
are given to be
ea () = a T
ea (eb ) = ba
(3.1.3a)
(3.1.3b)
This clearly gives N real-valued functions which satisfy Eq(3.1.3b). Now we will
show that they are linear and they constitute a basis for T .
The linearity of the real-valued functions is shown by the restriction
ea ( + ) = a + a
, R , T
(3.1.4)
To prove that {ea } is a basis for any given T we can define N real numbers,
denoted a by (ea ) = a . Then
() = (a ea ) = a (ea )
= a a
= a ea () T
(2.1.5)
(x y)
F
M
kx yk
|F (x) F (y)| M kx yk;
line 2 follows from linearity of the linear functional. Let
that F is continuous.
(2.1.6)
M
then it is clear
Lemma 3.3. Let E be the set of all continuous linear functionals on the normed
space (E, k k) and F E . Then kF k = supxE,kxk1 |F (x)| is a norm on E .
Proof. We simply need to check if the norm satisfies the following three conditions:
1. kF k > 0 if F 6= 0;
2. kF k = ||kF k C F E
3. kF + F 0 k kF k + kF 0 k F, F 0 E
Clearly conditions 1 and 2 are easily shown by definition of the norm. Condition
3 is satisfied by the triangle inequality.
Theorem 3.4. The set E of all continuous linear functionals on the normed space
(E, k k) is itself a Banach space with respect to pointless algebraic operations and
norm
kF k = supxE,kxk1 F (x)
(3.1.7)
Proof. E is a vector space over the same field as E. From Theorem 2.3, we know
that kF k is a real number and that it is finite. Lemma 2.3 also tells us that the given
norm is a norm on E . We will now show that E is complete.
Let (Fn ) be a Cauchy sequence in E so that kFn Fm k 0 as n, m .
Therefore x E |Fn (x) |Fm | 0 as n, m . We see that this is Cauchy
sequence of scalars; we denote the limit of this scalar sequence by F (x). We must
now show that F E and that Fn F with respect to the norm on E .
Let > 0 and pick n0 N such that m, n n0 kFm Fn k <
x E such that kxk 1 and m, n n0 ,
|Fm (x) Fn (x)| <
(3.1.8)
(3.1.9)
x
kxk
(3.1.10)
x
F
kF k
kxk
|F (x)|
kF k
kxk
(2.1.11)
(2.1.12)
y M
(3.1.13)
(3.1.14)
(3.1.15)
(2.1.16)
The last line comes from z C z + z = 2Re{z}. Eq(2.1.16) holds for any
C, it also holds for = tz where t > 0 and z C, where |z| = 1, then
z(x, y) = |(x, y)|. Thus
2t|(x, y)| + t2 kyk2 0
1
k(x, y)| tkyk2 .
2
(2.1.17)
(3.1.18)
(2.1.19)
By Lemma 2.7, z M .
Theorem 3.9. Riesz-Frechet Theorem
Let H be a Hilbert space and let F be a continuous linear functional on H. There
exists a unique y H such that
F (x) = (x, y)
Furthermore kyk = kF k.
x H.
(3.1.20)
(3.1.21)
(3.1.22)
(3.1.23)
z
kzk2
x H,
(3.1.24)
we have
(x, y) = F (x)
(3.1.25)
y
,
kyk
(3.1.26)
|F (y)|
|(y, y)|
=
= kyk.
kyk
kyk
(3.1.27)
Therefore kF k = kyk.
3.2
Tensor Product
Tensor products came around for vector spaces due to its inherent need in physics
and engineering. The following description of tensor products is of vector spaces.
Let R be a commutative ring and M and N be R-modules. We assume that
the rings R will have a multiplicative identity and that the modules are to be unital
1 m = m, m M . The product operation M R N , is calle the tensor product.
However, we will begin with the tensor product of vector spaces first.
Let V and W be vector spaces over a field K, and let {ei } and {fj } be a basis
for V and W respectively. Then the tensor product of these two vector spaces is
V K W defined to be the K-vector space with a basis of formal symbols ei fj ;
simply claim that these new symbols are linearly dependent by definition. Then the
tensors
or elements of the K-vector space are written in formal sums with the form
P
i,j cij ei fj with cij K. Furthermore, v V, w W we define v w to be the
element of V K W given by writing v and w in terms of the original bases of both
V and W then expanding as if the is a non commutative product. For example we
could let V = W = R2 Then we would have a 4 dimensional space R2 R R2 with 4
basis vectors: e1 e1 , e1 e2 , e2 e1 , e2 e2 . If we let v = e1 + 2e2 and w = 3e1 + e2 ,
then
v w = (e1 + 2e2 ) (3e1 + e2 ) = 3e1 e1 + e1 e2 6e2 e1 + 2e3 e3 (3.2.1)
If one were to pick another basis of R2 for the tensor product, we realize that v w
has a meaning in the space R2 R R2 independent of the chosen basis. Looking back
at a more generalized version, with modules, but the properties of the product are
to be satisfied in general, and not just on a basis: for R-modules M and N , their
tensor product M R N is an R-modules spanned, by a spanning set, by all symbols
m n, m M, n N . The symbols satisfy these distributive laws:
(m + m0 ) n = m n + m0 n
m (n + n0 ) = m n + m n0
r(m n) = (rm) n = m (rn)
r Rm, m0 M, n, n0 N
(3.2.2a)
(3.2.2b)
(3.2.2c)
(3.2.3a)
Since we know from above that ri (mi ni ) = (ri mi ) ni then let ri mi be renamed
just as mi . Therefore the above linear combination can be written as a sum
k
X
mi ni = m1 n1 + ... + mk nk .
(3.2.4)
i=1
The idea of two sums equal in M R N is not trivial in terms of the given description
of a tensor product. However, there is one case: let M and N be free R-modules with
bases {ei } and {fj } respectively. The resulting space M r N is then also a free module
with basis P
{ei fj }. An element is then equal to another when the coefficients of the
finite sum i,j cij ei fj are the same.
In a general case, where both M and N dont have bases, equality in M R N
relies on using a universal mapping property of the tensor product: M R N is the
universal object that turns bilinear maps on M N into linear maps. Before we
continue to define the tensor product rigorously, we will first define a bilinear map.
Definition 20. A function B : M N P , where M, N , and P are R-modules, is
called bilinear when it is linear in each argument with the other one fixed:
B(m1 + m2 , n) = B(m1 , n) + B(m2 , n)
B(rm, n) = rB(m, n)
B(m, n1 + n2 ) + B(m, n1 ) + B(m, n2 )
B(m, rn) = rB(m, n)
r R, m, m1 , m2 M, n, n1 , n2 N
(3.2.5a)
(3.2.5b)
(3.2.5c)
(3.2.5d)
M N
P
linear
bilinear
Q
B
map M N
M R N such that for any bilinear map M N
P there is a linear
L
map M R N
P making the following diagram commute:
M N
B
M R N
L
Despite the definition M R N is not explicitly given but instead gives a universal
mapping property involving it. Note that maps out of M N are bilinear and those
out of M R N are linear; bilinear maps out of M N turn into linear maps out of
M R N . We can show that any of the two constructions are the same.
Let T, T 0 be R-modules, and b : M N T and b0 : M N T 0 satisfy the
universal mapping property of the tensor product of M and N . From the definition,
the universality of b the map b0 factors uniquely through T : a unique linear map
f : T T 0 . Furthermore, the universality of b0 induces a unique factorization in B
through T 0 : a unique linear map f 0 : T 0 T . Combining the two together we have:
T
b
M N
f
0
T0
f0
b
T
(3.2.8)
(3.2.9a)
The other relations of a bilinear function also hold. This shows that the function
: M N M R N is bilinear. Now, we need to show that all bilinear maps out of
M N factor uniquely through the bilinear map . Suppose P is an R-module and
B : M N P is a bilinear map. Treat M N as a set, so B is just a function on
this set, the universal mapping property of free modules extends B from a function
M N toP to a linear function ` : FR (M N ) P with `((m,n) = B(m, n). We
want to show that ` make sense as a function on M R N ; show that D ker(`)
From the linearity of B,
B(m + m0 , n) = B(m, n) + B(m, n0 ), B(m, n + n0 ) = B(m, n) + B(m, n0 )
rB(m, n) = B(rm, n) = B(m, rbn)
(3.2.10a)
therefore
`((m+m0 ,n) ) = `((m,n) ) + `((m0 ,n) ), `((m,n+n0 ) ) = `((m,n) ) + `((m,n0 ) )
r`((m,n) ) = `((rm,n) ) = `((m,rn) ).
(3.2.11a)
Since ` is linear, these conditions must hold as well:
`((m+m0 ,n) ) = `((m,n) + (m0 ,n) ), `((m,n+n0 ) ) = `((m,n) + (m,n0 ) ),
`(r(m,n) ) = `((rm,n) ) = `((m,rn) ).
(3.2.12a)
Therefore the kernel of ` contains all the generators of the submodule D, so `
will induce a linear map L : FR (M N )/D P where L((m,n) + D) = `((m,n) ) =
B(m, n). Since FR (M N )/D = M R N and (m,n) + D = m n there for the
diagram below commutes as such:
M N
B
M R N
L
This diagram shows that every linear map B out of M N comes from a linear
map L out of M R N such that L(m n) = B(m, n), m M, n N . Now we
show that the linear map L is the only one that makes the diagram commute.
The definition of M R N as a quotient of the free module FR (M N ) tells us
that every element of FR (M N ) is a finite sum
r1 (m1 ,n1 ) + r2 (m1 ,n2 ) + + rk2 (mk ,nk ) .
(3.2.13)
(3.2.14)
where ai = bj = 0 for all but finitely many i and j. Since is bilinear we have
X
X
X
mn=
ai x i
bj y j =
ai b j x i y j
(3.2.16)
i
i,j
(3.2.17)
3.3
Now we will look at specific examples of tensor products limited to Hilbert spaces.
In defining the (Hilbert) tensor product H of two Hilbert spaces H1 , H2 , the
approach we take will utilize and emphasize what we did above, the universal
property of the tensor product. In terms of operators, rather than mappings, The
Hilbert space H is characterized, up to isomorphism, by the existence of a bilinear
mapping p : H1 H2 H. It has the following property that each suitable bilinear
mapping L from H1 H2 into a Hilbert space K has a unique factorization L = T p,
with T being a bounded linear operator from H into K.
Before the formal construction of the theory, we need to understand the intuitive
aspects of it. When x1 H1 , x2 H2 , we want to view the element p(x1 , x2 ) H as
a product x1 x2 . The linear combinations of such products form an everywheredense subspace of H. The bilinearity of p implies that these products satisfy certain
linear relations:
(x1 + y1 ) (x2 + y2 ) x1 x2 x1 y2 y1 x2 y1 y2 = 0
x1 , y1 H1 , x2 , y2 H2
(3.3.1a)
All the linear relations satisfied by product vectors can be achieved by use of the
bilinearity of p. The inner product on H satisfies the conditions:
hx1 x2 , y1 y2 i = hx1 , y1 ihx2 , y2 i
kx1 x2 k2 = hx1 x2 , x1 x2 i
= hx1 , x1 ihx2 , x2 i = kx1 k2 kx2 k2
kx1 x2 k = kx1 kkx2 k.
(3.3.2a)
(3.3.2b)
x1 H1 , ..., xn Hn .
(3.3.3)
If this is so then the lest such constant c is denoted by the norm of the mapping
kk. Then is a continuous mapping from H1 Hn C relative to the product
tor the norm topologies on the Hilbert spaces.
Theorem 3.13. Suppose that H1 , ..., Hn are Hilbert spaces and is a bounded multilinear functional on H1 Hn .
(i) The sum
X
X
yn Yn
has the same finite or infinite value for all orthonormal bases Y1 of H1 ,..., Yn of
Hn .
(ii) If K1 , ..., Kn are Hilbert spaces, Am B(Hm , Km ), (m = 1, ..., n), is a
bounded multilinear functional on K1 Kn , and
(x1 , ..., xn ) = (A1 x1 , ..., An xn )
x1 H1 , ..., xn Hn
(3.3.5)
then
X
y1 Y1
yn Yn
z1 Z1
zn Zn
yn Yn
z1 Z1
(3.3.7)
zn Zn
Therefore,
zn Zn
X
y1 Y1
yn Yn
y1 Y1
y1 Y1
X
yn Yn
z1 Z1
yn Yn z1 Z1
|(
zn Zn
zn Zn
z1 Z1
zn Zn
z1 Z1
zn Zn
(3.3.9a)
The third line is arrived by noticing that we can rewrite the zm Zm basis in
terms of ym Ym
X
X
z1 =
hz1 , y1 iy1 , ..., zn =
hzn , yn iyn .
(3.3.10)
y1 Y1
yn Yn
hzm , ym iym i
ym Ym
ym Ym
hzm , ym iym ,
ym Ym
ym Ym
|hym , zm i|2
(3.3.11a)
ym Ym
Using the same argument and exchanging the orthonormal bases we can show
equality.
Then for the proof of (ii), we suppose that 1 m n we choose and fix vectors
y1 Y1 , ..., ym1 Ym1 , zm+1 Zm+1 , ..., zn Zm . The mapping
z (A1 y1 , ..., Am1 ym1 , z, zm+1 , ..., zn ) : Km C
(3.3.12)
(3.3.13)
ym YM
|hAm ym , wi|2 =
ym YM
|hym , Am wi|2
ym YM
|hzm , wii|2
zm Zm
= kAm k2
(3.3.14a)
zm Zm
kAm k
ym Ym zm+1 Zm+1
y1 Y1
zn Zn
ym1 Ym1 zm Zm
zn Zn
(3.3.15)
Therefore
X
y1 Y1
kAn k2
y1 Y1
X
y1 Y1
y1 Y1
yn Yn
yn Yn
yn1 Yn1 zn Zn
kAn1 k2 kAn k2
X
X
|(A1 y1 , ..., An2 yn2 , zn1 , zn )|2
kA1 k2 kAn k2
X
z1 Z1
(3.3.16)
zn Zn
y1 Y1
kk2 =
X
y1 Y1
(3.3.17)
(3.3.18)
yn Yn
21
,
(3.3.19)
yn Yn
(3.3.20)
(3.3.21)
(3.3.22)
(3.3.23)
(3.3.24)
If U = 0, then
(3.3.25)
(3.3.26)
yY1 Yn
y1 Y1
yn Yn
X
y1 Y1
21
|f (y1 , ..., yn )|
yn Yn
X
21
X
2
2
|hx1 , y1 i| |hxn , yn i|
y1 Y1
= kf k
X
yn Yn
2
21
|hx1 , y1 i|
y1 Y1
X
21
|hxn , yn i|
yn Yn
= kf kkx1 k kxn k
From this, the equation
X
X
(x1 , ..., xn ) =
(3.3.27a)
(3.3.28)
yn Yn
(3.3.29)
(3.3.30)
y1 Y1
X
y1 Y1
X
yn Yn
yn Yn
2
|hy1 , v(1)i|
y1 Y1
X
|hyn , v(n)i|
yn Yn
= kv(1)k2 kv(n)k2 .
(3.3.31)
y1 Y1
X
y1 Y1
yn Yn
y1 Y1
yn Yn
X
hw(1), y1 ihy1 , v(1)i
hw(n), yn ihyn , v(n)i
yn Yn
(3.3.32)
(3.3.33)
(3.3.34)
is a basis of HSF.
Let us introduce the notion of the conjugate of a Hilbert space H. For a normal Hilbert space, H has the algebraic structure and inner product defined by the
mappings
(x, y) x + y : H H H
(a, x) ax : C H H
(x, y) hx, yi : H H C.
(3.3.35a)
(3.3.35b)
(3.3.35c)
(3.3.36a)
(3.3.36b)
(3.3.36c)
where
ax = a
x
(3.3.37)
is H.
It is also very easy to see that the conjugate Hilbert space of H
A useful aspect that arises is a subset of a Hilbert space is linearly independent,
orthogonal, or orthonormal, or an orthonormal basis of that space, if and only if it
has the same property relative to the conjugate Hilbert space. If H1 and H2 are
Hilbert spaces and T is a mapping from the set H1 into the set H2 , the linearity of
1 H
2 , and corresponds to conjugateT : H1 H2 is equivalent to linearity of T : H
2 and of T : H
1 H2 . Of course, continuity of T is the same
linearity of T : H1 H
in all four situations, when T is linear the operators have the same bound, since the
j .
norm on Hj is the same as that on H
Definition 24. Suppose that H1 , ..., Hn and K are Hilbert spaces and L is a mapping
from H1 Hn into K. L is a bounded multilinear mapping if it is linear
in each of its variables and there is a real number c such that
kL(x1 , ..., xn )k ckx1 k kxn k.
The least such constant c is denoted by kLk.
(3.3.38)
(3.3.39)
is a Hilbert-Schmidt functional on H1 Hn .
(ii) There is a real number d such that kLu k2 dkuk for each u K.
When these conditions are fulfilled, the least possible value of the constant d is
denoted by kLk2 .
A bounded multilinear mapping L : H1 Hn K is (jointly) continuous
relative to the norm topologies on the Hilbert spaces. We see that condition (ii)
actually flows from (i) by an application of the closed graph theorem to the mapping.
Theorem 3.15. Suppose that H1 , ..., Hn are Hilbert spaces.
(i) There is Hilbert space H and weak Hilbert-Schmidt mapping p : H1
Hn H with the following property: given any weak Hilbert-Schmidt mapping L
from H1 Hn into a Hilbert space K, there is a unique bounded linear mapping
T from H into K, such that L = T p; moreover, kT k = kLk2 .
(ii) If H0 and p0 have the properties attributed in (i) to H and p, there is a unitary
transformation U from H onto H0 such that p0 = U p.
(iii) If vm , wm Hm and Ym is an orthonormal basis of Hm , m = 1, ..., n, then
hp(v1 , ..., vn ), p(w1 , ..., wn )i = hv1 , w1 i hvn , wn i,
(3.3.40)
(3.3.41)
(3.3.42a)
(3.3.42b)
(3.3.43)
With y(1) Y1 , ..., y(n) Yn , orthonormality of the bases implies that y(1),...,y(n)
takes the value 1 at (y(1), ..., y(n)) and 0 elsewhere on Y1 Yn . Thus
p (y(1), ..., y(n)) = hp(y(1), ..., y(n)), i = hy(1),...,y(n) , i
X
X
=
yn Yn
y(1)Y1
(3.3.44a)
(3.3.44b)
y(n)Yn
(y1 ,...,yn )F
(y1 ,...,yn )F
21
|(y1 , ..., yn )|
(y1 ,...,yn )F
21
(y1 ,...,yn )F
kLu k2
21
|(y1 , ..., yn )|
(y1 ,...,yn )F
kukkLk2
X
(y1 ,...,yn )F
Hence
|(y1 , ..., yn )|
21
(3.3.45)
(y1 ,...,yn )F
kLk2
21
|(y1 , ..., yn )|
(3.3.46)
(y1 ,...,yn )F
Since
X
y1 Y1
(3.3.47)
yn Yn
it then follows from Eq(3.3.46) and the Cauchy criterion that the, unordered, sum
X
X
yn Yn
y1 Y1
yn Yn
(3.3.49)
Both L and T p are bounded and multiline and Ym has closed linear span, Hm , m =
1, ..., n then it follows that L = T p.
The condition that T p = L uniquely determines the bounded linear operator T ,
because the range of p contains the orthonormal basis p(Y1 Yn ) of H. u K,
Parsevals equation gives
kLu k22 =
y1 Y1
=
=
X
yn Yn
y1 Y1
yn Yn
y1 Y1
yn Yn
= kT uk2 kT k2 kuk2 ;
so we have kLk2 kT k, and thus kLk2 = kT k.
(3.3.50)
(3.3.51)
(3.3.52)
furthermore,
(3.3.53)
x1 H1 , ..., xn Hn , (3.3.54)
while
[p(H1 Hn )] = H
(3.3.55)
x H
(3.3.56)
(4.0.57)
Proof. Consider the algebraic tensor product A H, of the unital C -algebra and
the Hilbert Space H, and define a symmetric bilinear function h, i on this space by
setting
ha x, b yi = h(b a)x, yiH ,
(4.0.58)
and extending linearly, where h, iH is the inner product on H. From the definition,
is completely positive thereby ensuring that h, i is positive semidefinite since
X
n
j=1
aj x j ,
n
X
ai x i
i=1
x1
x1
.. ..
= n ((a a)) . , .
0,
H(n)
xn
xn
(4.0.59)
where h, iH(n) denotes the inner product not the direct sum H(n) of n copies of H,
given by
y1
x1
.. ..
= hx1 , y1 iH + + hxn , yn iH
(4.0.60)
. , .
H(n)
yn
xn
A result of positive semidefinite bilinear forms is that they satisfy the CauchySchwarz inequality,
|hu, vi|2 hu, ui hv, vi
(4.0.61)
(4.0.62)
(4.0.63)
will be an inner product. We let K denote the Hilbert space that is the completion
of the inner product space A H/N .
If a A, define a linear map (a) : A H A H by
X
X
(a)
ai x i =
(aai ) xi
(4.0.64)
A matrix factorization shows that the following inequality in Mn (A)+ is satisfied:
(ai a aaj ) ka ak (ai aj ),
(4.0.65)
and consequently,
X
X
(a)
aj xj , (a)
ai x i
X
X
=
h(ai a aaj )xj , xi iH ka ak
h(ai aj )xj , xi iH
i,j
i,j
2
= kak
X
aj x j ,
ai x i .
(0.0.12)
Therefore, (a) leave N invariant and consequently induces a quotient linear transformation on A H/N , which we will still denote by (a). The above inequality also
shows that (a) is bounded with k(a)k kak. Thus, (a) extends to a bounded
linear operator on K, which we will still denote by (a). It is straightforward to verify
that the map : A B(K) is a unital - homomorphism.
Now define V : H K via
V (x) = 1 x + N
(4.0.67)
(4.0.68)
x, y H,
(4.0.69)