Quantizing Analysis

download Quantizing Analysis

of 37

Transcript of Quantizing Analysis

  • 8/9/2019 Quantizing Analysis

    1/37

    Quantization of AnalysisKelvin Lui

    12/11/2014

    Department of MathematicsColby College

    2015

  • 8/9/2019 Quantizing Analysis

    2/37

  • 8/9/2019 Quantizing Analysis

    3/37

    Definition 8.   A  ring  is a datum  (R, +, ×, 0, 1)  where  R  is a set,  1, 0 ∈ R  and  +, ×are binary operations on  R  defined such that 

    1.   R  is an abelian group under  +  with identity element  0;

    2. The binary operator  ×   is associative and  1 × x =  x × 1 = x   ∀x ∈ R;

    3. Multiplication distributes over addition:

    x × (y + z ) = (x × y) + (x × z )

    (x + y) × z  = (x × z ) + (y × z )   ∀x,y,z  ∈ R

    If the operation   ×   is commutative,   x ×  y   =   y  × x   ∀x, y   ∈   R, we say   R   is acommutative ring. In addition, if 1 = 0 then the ring has only one element and isreferred to as the ”zero ring”. An example of a ring would be the integers  Z  or the

    integers modulo  n  Z/nZ  , n ∈  Z. If  R   is a ring, we can construct a new ring  R[t] of polynomials in t  with coefficients in  R.

    Definition 9.   If  R   is a ring, a subset  S  ⊆  R   is said to be a   subring   if it inherits the structure of a ring from  R, therefore  0, 1 ∈  S   and  S  is closed under the addition and multiplication operations in  R;  (S, +)  is a subgroup of  (R, +).

    An example would be the integers  Z are a subring of  Q.

    Lemma 2.1.  Subring Criterion Let  R  be a ring and  S  ⊆ R, then  S  is a sunning if and only if  1 ∈ S  and ∀s1, s2 ∈ S 

    we have  s1s2, s1 − s2 ∈ S .

    Proof.   (⇒) If we assume that  S   is a subring the proof is trivial.(⇐) Note  S  =  {0}  because 1  ∈  S . If  s1s2, s1 − s2  ∈  S,   ∀s1, s2  ∈  S  that implies

    that S  is an additive subgroup by the subgroup test. It is also easy to see that 0 ∈ S and that the other conditions for a subring hold.

    Definition 10.   A map   f   :   R   →   S  between rings   R   and   S   is said to be a   (ring)homomorphism  if 

    1.   f (1R) = 1S 

    2.   f (r1 +R r2) = f (r1) +S  f (r2)

    3.   f (r1.Rr2) = f (r1).S f (r2)

  • 8/9/2019 Quantizing Analysis

    4/37

    The image of a ring homomorphism  f   : R  →  S  is a subring of  S . If the image isall of  S  we say  f   is surjective, and if  f   :  R  → S   is an isomorphism then  ∃g   :  S toRsuch that  f  ◦ g  =  idS   and g ◦ f  = idR.

    We can also look at polynomials with coefficients in a ring. If  R  is a ring then the

    set  R[t] of sequences (an)n∈N, an ∈ R  where ∃N  ∈  N  such that  an  = 0, ∀n ≥  N . Wecan denote elements of  R[t] as

    N n=0

    antn.   (2.1.2)

    We will define addition and multiplication to be

    n=0ant

    n +M 

    n=0bnt

    n =

    max{N,M }

    n=0(an + bn)t

    n (2.1.3a)

    N n=0

    antn ×

    M n=0

    bntn =

    N +M n=0

    nk=0

    akbn−k

    tn (2.1.3b)

    and with this we see that  R[t] is a ring. The set of all sequences (an)n∈N  forms aring as well with the same definitions of addition and multiplication. It is called thering of formal power series in  t  and denoted as  R[[t]]. We view elements of  R[[t]] as”infinite sums” ,

     ∞n+0 ant

    n.

    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)

    and the  image  of  f   is 

    im (f ) = {s ∈ S   : ∃r ∈ R, f (r) = 2}.   (2.1.5)

    The image of a homomorphism is a subring of the target ring. The kernel, howeveris not a subring of the target ring. The kernel is closed under addition and multi-

    plication and since 0.x  = 0,   ∀x, it obeys a stronger kind of closure with respect tomultiplication. If  x  ∈  ker(f ) and  ∀r  ∈  R  then  f (x.r) =  f (x).f (r) = 0.f (r) = 0 sothat x.r ∈ ker(f ).

    Definition 13.  Let  R  be a ring. A subset  I  ⊆ R is called an  ideal  if it is a subgroupof  (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.

  • 8/9/2019 Quantizing Analysis

    5/37

  • 8/9/2019 Quantizing Analysis

    6/37

    C 1 + C 2 = {c1 + c2  :  c1 ∈ C i, i = 1, 2}   (2.2.1)

    C 1 +  C 2   is a subset of  R  and we claim it to be a single coset of  I . Definitely, it

    is a union of cosets, since  c1 + c2  ∼  r, for some  r  ∈  R  we have  r  = (c1 +  c2) +  i  forsome   i ∈  I  and so  r  = (c1 +  i) +  c2  ∈  C 1 +  C 2; the sum is an element of the unionof cosets. To show that it is a single coset, that addition is well-defined, fix elementsdi ∈ C i, i = 1, 2. Then if  ci ∈ C i  we have  ci =  di + ki, ki ∈ I , hence

    c1 + c2 + (d1 + k1) + (d2 + k2) = (d1 + d2) + (k1 + k2) ∈ (d1 + d2) + I.   (2.2.2)

    To give R/I  a ring structure, we need to define a multiplication operation. Giventwo cosets  C 1, C 2  define:

    C 1 · C 2 = {c1 · c2 :  ci ∈ C i, i = 1, 2}.   (2.2.3)We also claim that the set  C 1 · C 2   is contained in a single coset of  R/I , that is

    C 1 · C 2 + 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)   ∈I 

    ∈ d1 · d2 + I   (2.2.4)

    We get a binary operation on  R/I  by setting:

    C 1 × C 2 =  C 1 · C 2 + I    (2.2.5)

    Theorem 2.4.  The datum  (R/I, +, ×, 0 + I , i + I )  defines a ring structure on  R/I and moreover the map   1 :   R   →   R/I   given by   q (r) =   r  +  I   is a surjective ring homomorphism.

    Proof.   It is clear that   R/I   is an abelian group under + with an identity element0. We then check to see that the binary operation   ×   is associative and that themultiplication distributes over addition:

    C 1 × (C 2 × C 3) = (a1 + I ) × ((a2 + I ) × (a3 + I ))

    = (a1 + I )((a1 · a2 + I )) = (a1)(a2 · a3) + I = (a1 · a2)a3 + I  = (a1 · a2 + I ) × (a3 + I )

    = ((a1 + I ) × (a2 + I )) × (a3 + I )

    = (C 1 × C 2) × C 3

    ∀C 1, C 2, C 3 ∈ R/I,a1, a2, a3 ∈ R   (2.2.6)

  • 8/9/2019 Quantizing Analysis

    7/37

    a simple adaptation with the multiplicative identity will show associativity of 1R/O × C 1 =  C 1  =  C 1 × 1R/O, where 1R/O  = 1 + I .

    C 1 × (C 2 + C 3) = (a1 + I ) × ((a2 + I ) + (a3 + I ))= (a1 + I ) × ((a2 + a3) + I ) = a1(a2 + a3) + I 

    = (a1a2 + a1a + 3) + I 

    = (a1a2 + I ) + (a1a3 + I )

    = C 1 × C 2 + C 1 × C 3

    ∀C 1, C 2, C 3 ∈ R/I,a1, a2, a3 ∈ R   (2.2.7)

    A similar process can be done to show (C 1 + C 2) × C 3 = (C 1 × C 3) + (C 2 × C 3).Finally the map  q   :  R  →  R/I   is surjective and that it is a homomorphism can

    been seen immediately from the definitions.

    The map  q  :  R → R/I   is called the quotient homomorphism.

    Corollary 2.4.1.  Any ideal is the kernel of a ring homomorphism.

    Proof.   If  I   is an ideal and  q  :  R → R/I  is the quotient map then clearly ker(q ) = I .

    Next, we will look at comparisons between ideals in a quotient ring with ideals inthe original ring.

    Lemma 2.5.   Let  R  be a ring,  I   an ideal in  R   and  q   :  R →  R/I   the quotient map.If   J   is an ideal then   q (J )   is an ideal in   R/I , and if   K   is an ideal in   R/I   then q −1(K ) =  {r  ∈  R   :  q (r)  ∈  K }   is an ideal in  R   which contains   I . Moreover, these correspondences give a bijection between the ideals in  R/I  and the ideals in  R  which contain  I .

    Proof.   Let   q   :   R   →   R/I   denote the quotient homomorphism. If   J   is an ideal inR   we need to check that q(J) is an ideal in   R/I .   q   is a homomorphism so it is asubgroup of  R/I   under addition, we need to check the multiplicative property. Anyelement of   q (J ) has the form   q ( j),  ∀ j   ∈   J . From Theorem 2.4 we know that themap  q   is surjective and so for any element of  R/I  has the form  q (r), ∀r  ∈ R. Since

    q (r) · q ( j) = q (r · j) and  r · j ∈ J  j then q (J ) is an ideal in  R/I .If   K   ⊆   R/I   is an ideal then   q −1(K ) is an ideal of   R. Let   t1, t2   ∈   q −

    1(K ) theq (t1  + t2) =   q (t1) + q (t2)   ∈   K   because   K   is an ideal, then   t1  + t2   ∈   q −1(K ). If  t  ∈  q −1(K ) and  r  ∈  R  then  q (rṫ) =  q (r) · q (t)  ∈  K   because  K   is an ideal of  R/I .Then this means that  r · t ∈ q −1(K ) and therefore  q −1(K ) is an ideal in R.

    Note that  q (q −1(K )) = K , true for surjective map  q  of sets. If  J   is an ideal of  Rthen we claim that the composition  q −1(q (J )) is the ideal  J  + I . Assuming that if 

  • 8/9/2019 Quantizing Analysis

    8/37

    I  ⊆ J   then q −1(q (J )) =  J , hence  q   is bijective when restricted to ideals containing  I which is required. Now we need to check that  q −1(q (J )) + J  +  I . Let  x ∈  q −1(q (J ))then   q (x) =   q ( j) for some   j   ∈   J   and so   q (x −  j ) = 0, then   x −  j   ∈   I , becauseq (J ) ∈ R/I . However, x  =  j + (x − j) ∈ J  + I  so that q −1(q (J )) ⊆ J  + I . The reverse

    inclusion is clear because q ( j + i) = ( j + i) + I  = j  + I,  ∀ j  ∈ J, i ∈ I .

    Theorem 2.6.  First Isomorphism theorem Let  f   : R  →  S  be a homomorphism of rings, and let  I  = ker (f ). Then  f   induces 

    an isomorphism  f̄   : R/I  → im (f )  given by:

    f̄ (r + I ) = f (r) (2.2.8)

    Proof.   Note that if  r − s ∈  I   then  f (r − s) = 0 and therefore  f (r) = f (s),  f   take asingle value on each coset  r + I  so  f̄   is well-defined. Clearly from the definition of the

    ring structure on R/I  it is a ring homomorphism. It is clearly surjective, if  s ∈ im(f )then s  =  f (r) for some r  ∈ R thus s  =  f̄ (r + I ). To see that  f̄  is injective it is enoughthat ker( f̄ ) = 0. If  f̄ (R + I ) = 0 ⇒ f (r) = 0 hence r  ∈ I  so that  r + I  = I .

    Let R be a ring and  A, B  be subsets of  R. Let  A + B = {a + b :  a ∈ A, b ∈ B}.

    Theorem 2.7.   Second Isomorphism theorem Let  R  be a ring and  A  a subring of  R, B  an ideal of  R. Then  A +B  is a subring of 

    R and the restriction of the quotient map  q  :  R → R/B   to A  induces an isomorphism 

    A/A ∩ B ∼= (A + B)/B   (2.2.9)Proof.   First let us check that  A + B   is a subring of  R. It is clear that 1  ∈  A + Bsince 1 ∈ A and for 1 ∈ A + B, 1 = 1 + 0. Then using the two other conditions of thesubring criterion

    (a1 + b1) × (a2 + b2) = a1a2 + b1a2 + a1b2 + b1b2   ∈B

    ∈ A + B   (2.2.10a)

    (a1 + b1) + (a2 + b2) = (a1 + a2) + (b1 + b2) ∈ A + B   (2.2.10b)

    It is easy to see that  A ∩ B   is an ideal in  A  and that  B   is an ideal in  A + B, sotherefore the two quotients  A/A ∩ B   and (A +  B)/B   exist. We let   q   :   R   →   R/Bbe the quotient map. The restriction of  q   to  p   :  A  →  R/B, whose image is clearly(A + B)/B, so by the first isomorphism theorem it is enough to check that the kernelof  p  is in A ∩ B. It is clear: if  a ∈ A has  p(a) = 0 then  a + B = 0 + B  so that  a ∈ Band so  a ∈ A ∩ B.

  • 8/9/2019 Quantizing Analysis

    9/37

    Theorem 2.8.  Universal property of Quotients If   f   :   R   →   S   is a homomorphism and   I   ⊆   ker (f )   is an ideal, then there is a 

    unique homomorphism  f̄   :  R/I   →  S   such that  f   =  f̄  ◦ q  where  q   :  R  →  R/I   is the quotient map.

    Proof.   Note that since  q   is surjective the requirement that  f̄ (q (r)) =  f (r) uniquelydetermines  f̄   if it exists. But if   r1  −  r2   ∈   I   then since   I   ⊆   ker(f ), we have 0 =f (r1 − r2) = f (r1) − f (r2) hence  f (r1) = f (r2). Follows that  f   is constant on the  I -cosets and so it induces a unique map  f̄   : R/I  → S  such that  f̄ ◦q  =  f . It is immediatefrom the definitions of the ring structure on  R/I   that  f̄  is a homomorphism.

    Theorem 2.9.  Third Isomorphism theorem If  I  ⊆ J  are ideals, then  J/I  = { j + I   : j  ∈ J }  is an ideal in  R/I   and 

    (R/I )/(J/I ) ∼= R/J    (2.2.11)

    Proof.   Let  q I   :  R  →  R/I   and let  q J   :  R  →  R/J  be the two quotient maps. By theuniversal property for the quotient q J   : R toR/J/ applied to the homomorphism q I  wesee that there is a homomorphism q̄ I   : R/J  → R/I  induced by the map q I   : R → R/I and q̄  ◦ q J   =   q I . Clearly q̄ I   is surjective since   q I , and if q̄ (r +  j) = 0 then sincer + J  = q J (r) so that q̄ I (r + J ) = q I (r) we have  r ∈  I  so that ker(q̄ I ) = J/I  and theresult follows from the first homomorphism theorem.

    2.3 Modules

    Definition 14.   Let  R   be a ring with identity   1R. A  module  over  R   is an abelian group (M, +)   together with a multiplication action of  R  on  M  satisfying:

    1R.m =  m,   ∀m ∈ M    (2.3.1a)

    (r1.r2)m =  r1(r2.m),   ∀r1, r2 ∈ R, m ∈ M    (2.3.1b)

    (r1 + r2)m =  r1.m + r2.m,   ∀r1, r2 ∈ R, m ∈ M    (2.3.1c)

    r.(m1 + m2) = r.m1 + r.m2,   ∀r ∈ R, m1, m2 ∈ M    (2.3.1d)

    A concrete sense of a module  M  over a ring R, could be quickly understood if one

    were to let  R  be a field, then the module would simply be a vector space over  R. Amodule over a field are essentially vector spaces over that field.

    Definition 15.   If  M   is an  R-module, a subset  N   ⊆ M   is called a  submodule   if it is an abelian subgroup of  M   and whenever  r ∈ R  and  n ∈ N   then  r.n ∈ N .

    One extension from linear algebra is that of linear independence.

  • 8/9/2019 Quantizing Analysis

    10/37

    Definition 16.  If  M  is a module over  R, we say a set  S  ⊆ M  is linear independent if whenever we have an equation  r1s1 + r2s2 + ... + rksk  = 0 for  ri ∈ R, s1 ∈ S  1 ≤ i ≤ kwe have  r1  =  r2  =  ...  =  rk  = 0. We can say that a set  S  is a basis for a module  M   if and only if it is linearly independent and it spans  M . Any module which has a basis 

    is called a   free module.

    On the other hand there is a notion of a spanning set.

    Definition 17.   A  spanning set  for an  R-module is a subset whose finite  R-linear combinations fill up the module. These always exist, since the entire module is a spanning set.

    To continue on we assume that all rings   R  are integral domains. As for vectorspaces, modules have a natural notion of a quotient module. An essential aspect of the quotient module is the condition that  N  be a submodule so that the multiplicationon M  induced a module structure on  M/N .

    Definition 18.   If  N   is a submodule of  M , then in particular it is a subgroup of an abelian group, so we can form the quotient   M/N . The module   M/N   is called the quotient module  of  M   by  N . We define multiplication on the quote module by:

    r.(m + N ) = r.m + N, rinR,m + N  ∈ M/N    (2.3.2)

    The multiplication well defined because if  m1 + N  = m2 + N  we then have  m1 −m2  ∈ N , same as quotient rings, and so  r.(m1 − m2) ∈  N   ⇒ r.m1 + N  = r.m2 + N .Similar to rings, we can define a natural analogue of linear maps.

    Definition 19.   If   M 1, M 2   are modules, we say that   φ   :   M 1   →   M 2   is a   modulehomomorphism  if:

    φ(m1 + m2) = φ(m1) + φ(m2),   ∀m1, m2 ∈ M 1   (2.3.3a)

    φ(r.m) = r.φ(m),   ∀r ∈ R, m ∈ M    (2.3.3b)

    Lemma 2.10.  Submodule correspondence Let  M   be an  R-module and  N   a submodule. Let  q   :  M   →  M/N  be the quotient 

    map. If   S   is a submodule of   M   then   q (s)   is a submodule of   M/N , while if   T   is a submodule of  M/N   then  q −1(T ) is a submodule of  M . Moreover, the map T   → q −1(T )gives an injective map from submodules of  M/N  to the submodules of  M  which contain N , thus submodules of  M/N  correspond bijectively to submodules of  M  which contain N .

    Proof.  We can check from the definitions that  q (S ) and q −1(T ) are submodules of  N and  M   respectively. As the proof for both are the same, we will only prove it forq (S ).

    Let   S   be a submodule of   M   and   q   :   M   →   M/N   be the quotient map. Lets1, s2   ∈   S   then   q (s1), q (s2)   ∈   M/N . Since   S   is a submodule then   s1  +  s2   ∈   S   ⇒

  • 8/9/2019 Quantizing Analysis

    11/37

    q (s1 +  s2) =  q (s1) +  q (s2)  ∈  M/N . Finally, if  r  ∈  R  then  r.s ∈  S , by definition of submodule, and φ(r.s) = r.φ(s) ∈ M/N . Therefore  q (S ) is a submodule of  M/N .

    Now, let   T   be any subset of   M/N . We have that   q (q −1(T )) =   T   is surjectivebecause  q  is surjective. Also because know that  q −1(T ) is always a submodule in  M ,

    the map of  S   →  q (S ) is a surjective map from submodules in  M   to submodules inM/N   and that   T   →   q −1(T ) is an injective map, furthermore since  q (N ) = 0   ∈   T for any submodule  T   of   M/N   we have  N   ⊆   q −1(T ) so that the image of the mapT   →   q −1(T ) consists of submodules of  M   which contain  N . We only now need tocheck that the submodules of  M  of the form  q −1(T ) are precisely these submodules.

    Suppose that   S   is an arbitrary submodule of   M   and consider   q −1(q (S )). Bydefinition

    q −1(q (S )) = {m ∈ M   : q (m) ∈ q (S )}

    = {m ∈ M   : ∃s ∈ S   m + N  = s + N }= {m ∈ M   : m ∈ s + N }   (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  φ  :  M  → N  is a homomorphism of  R-modules, and  S  is a submodule 

    of  M   with  S  ⊂  ker (φ). Then there is a unique homomorphism  φ̄   :  M/S   →  N   such 

    that  φ  =  φ̄◦q  where  q  :  M toM/S  is the quotient homomorphism, that is the following diagram commutes:

    M N 

    M/S 

    φ

    q    φ̄

    Moreover, ker (φ̄)  is the submodule ker (φ)/S  = {m + S  :  m ∈ ker (φ)}.

    Proof.   q  is surjective, so the formula  φ̄(q (m)) = φ(m) determines uniquely the valuesof  φ̄, thus φ̄ is unique if it exists. If  m−m ∈ S  then because by assumption S  ⊆ ker(φ)it follows that 0 = φ(m − m) = φ(m) − φ(m) since φ  is a homomorphism. Thereforeφ   is constant on the  S -cosets. This induces a map  φ̄(m +  S ) =  φ(m).   φ̄   is then ahomomorphism because it follows directly from the module structure on the quotientM/S , and by definition  φ =  φ̄ ◦ q .

  • 8/9/2019 Quantizing Analysis

    12/37

    φ̄

    (m1 + S ) + (m2 + S )

    = φ̄

    m1 + m2 + S 

    = φ(m1 + m2)

    = φ(m1) + φ(m2)

    = φ̄(m1 + S ) + φ̄(m2 + S ) (2.3.5a)

    φ̄

    r.(m + S )

    = φ̄(r.m + S )

    = φ(r.m) = r.φ(m)

    = r.φ̄(m + S ) (2.3.5b)

    Note that  φ̄(m + S ) = φ(m) = 0 if and only if  m ∈ ker(φ), therefore the kernel of φ̄  is m + S  ∈ ker(φ)/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  φ  :  M  → N  is a homomorphism then  φ induces an isomorphism  φ̄ :  M/ker (φ) →

    im (φ).

    Proof.   Let K  = ker(φ). Using the universal property of quotients to  K , then we haveker(φ̄) = ker(φ)/ker(φ) = 0 it follows that  φ̄   is injective and therefore induces anisomorphism onto its image which from the equation  φ̄ ◦ q  =  φ  = I’m(φ).

    Theorem 2.13.   Second isomorphism theorem If  M   is an  R-module and  N 1, N 2  are submodules of  M   then 

    (N 1 + N 2)/N 2 ∼= N 1/N 1 ∩ N 2   (2.3.6)

    Proof.   First, let  q   :  M   →  M/N 2  be the quotient map, which restricts  p  from  N 1   toM/N 2, and whose image is clearly (N 1 + N 2)/N 2. Therefore by the first isomorphismtheorem it is only necessary to check that the kernel of  p   is  N 1 ∩ N 2. If  n ∈ N 1  has

     p(n) = 0 then  n + N 2 = 0 + n2  so that  m ∈ N 2  and so n ∈ N 1 ∩ N 2.

    Theorem 2.14.  Third isomorphism theorem Suppose that  N 1 ⊆ N 2  are submodules of  M . Then we have 

    (M/N 1)/(N 2/N 1) ∼= M/N 2   (2.3.7)

  • 8/9/2019 Quantizing Analysis

    13/37

    Proof.   We first let   q i   :   M   →   M/N i   for   i   = 1, 2. Using the universal property of quotients for  q 1  we see that there is a homomorphism q̄ 2   :  M/N 1  →  M/N 2   inducedby the map   q 2   :  M   →   M/N 2, with the kernel being ker(q 2/N 1) =   N 2/N 1   and q̄ 2 ◦q 1  =  q 2. Therefore q̄ 2   is surjective because  q 2   is and the result follows from the first

    isomorphism theorem.

    These proofs mainly followed that of the proofs for the isomorphism theorems’ of rings.

    3 Tensors and Tensor Products

    3.1 Dual Spaces

    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 functionsdefined on a real vector space  T ,  f   :  T   →  R. The set of all such functions is thengiven the structure

    (f  +  g)(v) = f (v) + (v),   ∀v ∈ T    (3.1.1a)

    (αf )(v) = α(f (v))   ∀v ∈ T    (3.1.1b)

    0(v) = 0   ∀v ∈ T    (3.1.1c)

    (−f )(v) = −(f (v))   ∀v ∈ T,   (3.1.1d)

    which defines a vector space. Furthermore, we will restrict the space of all real-valued functions to only those that are linear

    f (αu + β v) = αf (u) + βf (v)   ∀α, β  ∈ R, ∀u, v ∈ T.   (3.1.2)

    The real-valued linear functions on a real vector space are also called linear func-tionals and the space constitutes a vector space called the dual of   T , denoted byT ∗.

    For simplicity we will point out the two types of vectors, those in  T  an those inT ∗, and distinguish them from one another. For the basis vectors of  T ∗ we will usesuperscripts and the components of vectors we will use subscripts, therefore if  {ea}

    is a basis of  T ∗, then a vector  λ ∈ T ∗ has the unique expression  λ  = a λaea.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 .

  • 8/9/2019 Quantizing Analysis

    14/37

    Proof.  First, let  T   be a  N -dimenionsal vector space and let  {ea}  be the given basisof   T . We define  {ea}   to be real-valued functions which maps   λ   ∈   T   into the realnumber  λa, which is the  ath component relative to the basis vectors of  T , {ea}. Theyare given to be

    ea(λ) = λa λ ∈ T    (3.1.3a)

    ea(eb) = δ ab   (3.1.3b)

    This clearly gives  N   real-valued functions which satisfy Eq(3.1.3b). Now we willshow that they are linear and they constitute a basis for  T ∗.

    The linearity of the real-valued functions is shown by the restriction

    ea(αλ + βµ) = αλa + βµa ∀α, β  ∈ R   ∀λ, µ ∈ T    (3.1.4)

    To prove that {ea

    }  is a basis for any given ν  ∈ T ∗

    we can define  N  real numbers,denoted ν a  by ν (ea) = ν a. Then

    ν (λ) = ν (λaea) = λaν (ea)

    = λaν a

    = ν aea(λ)   ∀λ ∈ T    (2.1.5)

    Therefore for any  ν   ∈  T ∗ we have  ν   =  ν aea, showing that  {ea}  stands  T ∗. Theindependence of the basis vectors comes from Eq(3.1.3b). Thus dim(T ∗) = dim(T ).

    Theorem 3.2.   Let   F   be a linear functional on a normed space   (E ,   · ). The   following statements are then equivalent to each other:

    1.   F   is continuous;

    2.   F  is continuous at 0;

    3. sup  {|F (x)| : x ∈ E, x ≤ 1} < ∞

    F   is continuous if and only if it is bounded on the unit ball of  E .

    Proof.   From statement 1 the proof of statement two is trivial.Assuming statement 2, we know  ∀ > 0  ∃δ > 0 such that  x < δ  ⇒ |F (x)| < .

    Take    = 1 then there exists a  δ >  0 such that  x  < δ   ⇒ |F (x)|  <   1. Therefore∀x  ∈  E  such that  x ≤  1 we have   δx

    2   < δ  ⇒ |F ( δx

    2 )|  

  • 8/9/2019 Quantizing Analysis

    15/37

    (x − y)

    x − y

    ≤ M 

    |F (x) − F (y)| ≤ M x − y; (2.1.6)

    line 2 follows from linearity of the linear functional. Let δ  ≤   M 

      then it is clearthat F   is continuous.

    Lemma 3.3.   Let   E ∗ be the set of all continuous linear functionals on the normed space  (E, · )  and  F  ∈ E ∗. Then  F  = supx∈E,x≤1|F (x)|  is a norm on  E 

    ∗.

    Proof.   We simply need to check if the norm satisfies the following three conditions:

    1.   F  > 0 if  F   = 0;

    2.   λF  = |λ|F  ∀λ ∈ C ∀F  ∈ E ∗

    3.   F  + F  ≤ F  + F  ∀F, F  ∈ E ∗

    Clearly conditions 1 and 2 are easily shown by definition of the norm. Condition3 is satisfied by the triangle inequality.

    Theorem 3.4.  The set  E ∗ of all continuous linear functionals on the normed space (E, · )   is itself a Banach space with respect to pointless algebraic operations and 

    norm 

    F  = supx∈E,x≤1F (x)   (3.1.7)

    Proof.   E ∗ is a vector space over the same field as  E . From Theorem 2.3, we knowthat F  is a real number and that it is finite. Lemma 2.3 also tells us that the givennorm is a norm on  E ∗. We will now show that  E   is complete.

    Let (F n) be a Cauchy sequence in   E ∗ so that   F n − F m →   0 as   n, m   → ∞.

    Therefore ∀x  ∈  E   ⇒ |F n(x) − |F m| →  0 as  n, m  → ∞. We see that this is Cauchysequence of scalars; we denote the limit of this scalar sequence by  F (x). We mustnow show that  F  ∈ E ∗ and that  F n → F  with respect to the norm on  E ∗.

    Let  > 0 and pick n0 ∈ N such that  m, n ≥ n0 ⇒ F m − F n < ∀x ∈ E  such that x ≤ 1 and m, n ≥ n0,

    |F m(x) − F n(x)| <   (3.1.8)

    Letting m → ∞, ∀n ≥ n0  and ∀x  x ≤ 1,

    |F (x) − F n(x)| ≤   (3.1.9)

  • 8/9/2019 Quantizing Analysis

    16/37

    Since  ∀x   ∈   E      x ≤   1 implies   F   − F 0   and   F n0   are bounded, then   F   is alsobounded on the unit ball and by Theorem 2.3 it is continuous. Thus   F   ∈   E ∗.Whenever  n ≥ n0 ⇒ F  − F n ≤  implies that  F n → F   ∈ E ∗ and so E ∗ is complete.

    Theorem 3.5.   For any vector   x   in a normed space   E   and any continuous linear  functional  F   on  E ,

    |F (x)| ≤ F x.   (3.1.10)

    Proof.  Suppose that  x ∈ E   x = 0. Then   xx

     is a unit vector and therefore

      x

    x

    ≤ F 

    |F (x)|x   ≤ F    (2.1.11)

    Lemma 3.6.   Let  V  be an inner product space.   ∀x,y,z  ∈ V , if  (x, z ) = (y, z )   ∀z  ∈V   ⇒ x =  y.

    Proof.   If (x, z ) = (y, z ) then

    0 = (x, z ) + (−1)(y, z )

    = (x, z ) + (−y, z )(x − y, z ) (2.1.12)

    If this is true for all  z  ∈ V   then z  =  x − y ⇒ x − y = 0.

    Lemma 3.7.   Let  M  be a linear subspace of an  H   and let  x ∈  H . Then  x ∈  M ⊥ if and only if 

    x − y ≥ x ∀y ∈ M    (3.1.13)

    Proof.   (⇒) If  x ∈ M ⊥

    then, ∀y ∈ M x and y are orthogonal and Pythagoras’ theoremgives

    x − y2 = x2 + y2 ≤ x2 (3.1.14)

    (⇐) Suppose that Eq(3.1.13) holds.   ∀y ∈ M   and  ∀λ ∈ C, then  λy  ∈ M   and

    x − λy2 ≥ x2.   (3.1.15)

  • 8/9/2019 Quantizing Analysis

    17/37

    Rewriting this as an inner product we have

    (x − λy,x − λy) ≥ x2

    x2 − λ̄(x, y) − λ(x, y)− + |λ|2y2 ≥ x2

    −2Reλ̄(x, y) + |λ|2y2 ≥ 0.   (2.1.16)

    The last line comes from  ∀z   ∈  C  ⇒  z  + z̄  = 2Re{z }. Eq(2.1.16) holds for anyλ   ∈   C, it also holds for   λ   =   tz   where   t >   0 and   z   ∈   C, where   |z |   = 1, thenz̄ (x, y) = |(x, y)|. Thus

    −2t|(x, y)| + t2y2 ≥ 0

    (x, y)| ≤ 1

    2

    ty2.   (2.1.17)

    Letting t → ∞  the result shows (x, y) = 0.

    Lemma 3.8.   Let  M  be a closed linear subspace of a Hilbert space  H  and let  x ∈ H .There exists  y ∈ M z  ∈ M ⊥ such that  x =  y  + z .

    Proof.   Take  y  ∈ M  and let it be the closest point to  x  in  M , by definition

    x − y ≤ x − m ∀m ∈ M.   (3.1.18)

    Let z  =  x − y  then  x  =  y  + z .   ∀m ∈ M  ⇒ y + m ∈ M  and so

    z  = x − y ≤ x − (y + m)

    ≤ z − m ∀m ∈ M.   (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)   ∀x ∈ H.   (3.1.20)

    Furthermore  y = F .

  • 8/9/2019 Quantizing Analysis

    18/37

    Proof.   y  is unique because

    (x, y) = F (x) = (x, y)   ∀x ∈ H    (3.1.21)

    implies y  =  y

    by Lemma 2.6.A trivial case of  F   is when F  is the zero operator and we take  y  = 0.Let M  be the kernel of the linear functional F ,

    M  = KerF   = {x ∈ H   : F (x) = 0},   (3.1.22)

    and  M   is a proper closed subspace of  H . From Lemma 2.8, we know that  H   =M ⊕ M ⊥ and therefore M ⊥ = {0}. Let z  ∈ M ⊥  z  = 0. We can scale z  with a scalarsuch that  F (z ) = 1. Then pick a  z  ∈ M ⊥ such that for any x ∈ H 

    x = (x − F (x)z ) + F (x)z    (3.1.23)

    Because  H   is a direct some of  M  and it’s orthogonal complement  M ⊥, the firstterm on the right hand expression is an element of   M   and the second term is anelement of  M ⊥. Taking an inner product of both sides with  z ,

    (x, z ) = (F (x)z, z ) = F (x)z 2 ∀x ∈ H,   (3.1.24)

    since z  ⊥ M . If we let  y  =   zz2

      we have

    (x, y) = F (x) (3.1.25)

    If  x ≤ 1 then by Cauchy-Schwarz,

    |F (x)| = |(x, y)| ≤ xy ≤ y   (3.1.26)

    Let x =   yy

    , which is a unit vector, and therefore

    F  ≥ |F (x)| =  |F (y)|

    y  =

      |(y, y)|

    y  = y.   (3.1.27)

    Therefore F  = y.

    3.2 Tensor Product

    Tensor products came around for vector spaces due to its inherent need in physicsand 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 thatthe rings  R  will have a multiplicative identity and that the modules are to be unital1 · 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.

  • 8/9/2019 Quantizing Analysis

    19/37

    Let  V   and  W   be vector spaces over a field  K , and let  {ei}  and  {f  j}  be a basisfor   V   and   W   respectively. Then the tensor product of these two vector spaces isV   ⊗K  W   defined to be the   K -vector space with a basis of formal symbols   ei ⊗ f  j;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 formi,j cijei ⊗ f  j  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 bothV   and W  then expanding as if the ⊗   is a non commutative product. For example wecould let  V   = W   =  R2 Then we would have a 4 dimensional space  R2 ⊗R R

    2 with 4basis 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 ⊗whas a meaning in the space  R2 ⊗R R2 independent of the chosen basis. Looking backat a more generalized version, with modules, but the properties of the product  ⊗  areto be satisfied in general, and not just on a basis: for R-modules  M   and   N , theirtensor product  M  ⊗R N   is an R-modules spanned, by a spanning set, by all symbolsm ⊗ n, m ∈ M, n ∈ N . The symbols satisfy these distributive laws:

    (m + m) ⊗ n =  m ⊗ n + m ⊗ n   (3.2.2a)

    m ⊗ (n + n) = m ⊗ n + m ⊗ n (3.2.2b)

    r(m ⊗ n) = (rm) ⊗ n =  m ⊗ (rn) (3.2.2c)

    ∀r ∈ Rm,m ∈ M,n,n ∈ N 

    In the space of  M  ⊗R N  there are usually more elements than products  m ⊗ n.The general element of  M ⊗R N , similar to linear algebra, is called a tensor and is anR-linear combination

    r1(m1 ⊗ n1) + r2(m2 ⊗ n2) + ... + rk(mk + nk),   (3.2.3a)

    k ≥ 1, ri ∈ R, mi ∈ M, ni ∈ N 

    Since we know from above that  ri(mi ⊗ ni) = (rimi) ⊗ ni then let rimi be renamed just as  mi. Therefore the above linear combination can be written as a sum

    ki=1

    mi ⊗ ni  =  m1 ⊗ n1 + ... + mk ⊗ nk.   (3.2.4)

    The idea of two sums equal in M ⊗RN  is not trivial in terms of the given descriptionof a tensor product. However, there is one case: let M  and N  be free R-modules with

  • 8/9/2019 Quantizing Analysis

    20/37

    bases {ei} and {f  j} respectively. The resulting space M ⊗rN  is then also a free modulewith basis {ei ⊗ f  j}. An element is then equal to another when the coefficients of thefinite sum

     i,j cijei ⊗ f  j  are the same.

    In a general case, where both  M   and  N  don’t have bases, equality in  M  ⊗R N 

    relies on using a universal mapping property of the tensor product:   M  ⊗R N   is theuniversal object that turns bilinear maps on   M  × N   into linear maps. Before wecontinue 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) (3.2.5a)

    B(rm,n) = rB(m, n) (3.2.5b)

    B(m, n1 + n2) + B(m, n1) + B(m, n2) (3.2.5c)

    B(m,rn) = rB(m, n) (3.2.5d)∀r ∈ R,m,m1, m2 ∈ M,n,n1, n2  ∈ N 

    This idea of bilinearity can be extended further to multilinearity.Now, any bilinear map  M  × N  → P  to an  R- module P  can be composed with a

    linear map P  → Q to induce a map  M  × N  → Q that is bilinear.

    M  × N P 

    Q

    bilinear

    bilinear   linear

    The construction of the tensor product of   M   and   N   will be the solution to auniversal mapping problem: find an  R-module T and bilinear map  b  : M  × N   →  T such that every bilinear map on  M  × N  is the composite of the bilinear map b  and aunique linear map out of  T . By definition then,

    Definition 21.  The  tensor product  M ⊗RN  is an  R-module equipped with a bilinear 

    map M × N   ⊗−→ M ⊗R N  such that for any bilinear map  M × N 

      B−→ P  there is a linear 

    map  M  ⊗R N   L−→ P   making the following diagram commute:

    M  × N M  ⊗R N 

    B   L

  • 8/9/2019 Quantizing Analysis

    21/37

    Despite the definition M ⊗R N  is not explicitly given but instead gives a universalmapping property involving it. Note that maps out of  M  × N  are bilinear and thoseout 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  be  R-modules, and   b   :  M  × N   →  T   and   b :  M  × N   →  T  satisfy theuniversal mapping property of the tensor product of  M   and  N . From the definition,the universality of   b   the map   b factors uniquely through   T : a unique linear mapf   :  T   →  T . Furthermore, the universality of  b induces a unique factorization in  Bthrough T : a unique linear map  f  : T  → T . Combining the two together we have:

    M  × N T 

    b

    b

    b

    We also notice that there is an unique mapping from   f  ◦ f   :   T   →   T . This isalso because of the universality of the map   b   to the   R-module   T . We can let thecomposition map of  f  ◦ f  = idT  be the identity map. Similarly we have  f ◦ f  = idT .Therefore both  T   and  T  are isomorphic R-modules by  f   and also  f  ◦ b  =  b, whichmeans f   identifies b  with  b. Any two tensor products of  M   and N  can be identified

    with each other in a unique way compatible with the distinguished bilinear maps tothem from  M  × N .

    Theorem 3.10.  The tensor product of  M   and  N  exists.

    Proof.   The tensor product of  M   and  N ,  M  ⊗R N  will be the quotient module of afree   R-module. For any set   S , we will write the free   R-module on   S   as   F R(S ) =

    s∈S  Rδ s, where  δ s  is the term with 1 in the  s-position and 0 everywhere else. ie. if s = s ∈  S  ⇒  rδ s − rδ s   has  r   in the  s-position,  −r in the  s-position, 0 elsewhere.Let S  = M  × N :

    F R(M  × N ) = (m,n)∈M ×N 

    Rδ (m,n).   (3.2.6)

    Let D  be a submodule of  F R(M  × N ) spanned by all the elements of:

    δ (m+m,n) − δ (m,n) − δ (m,n), δ (m,n+n) − δ (m,n) − δ (m,n), δ (rm,n) − δ (m,rn)

    rδ (m,n) − δ (rm,n), rδ (m,n) − δ (m,rn).   (3.2.7a)

  • 8/9/2019 Quantizing Analysis

    22/37

    The tensor product is then the quotient module by  D:

    M  ⊗R N   := F R(M  × N )/D.   (3.2.8)

    We write the coset δ (m,n)

    +D in M ⊗R

    N  as m⊗

    n. Remembering that the elementsof  D  are equivalent to 0 in  F R(M  × N )/D  gives us

    δ (m+m,n) ≡ δ (m,n) + δ (m,n)

    (m + m) ⊗ n =  m ⊗ n + m ⊗ n ∈ M  ⊗R N.   (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 andB :  M  × N  → P   is a bilinear map. Treat  M  × N  as a set, so  B   is just a function onthis set, the universal mapping property of free modules extends  B   from a function

    M  × NtoP   to a linear function     :  F R(M  × N )  →  P   with   (δ (m,n)   =  B(m, n). Wewant to show that    make sense as a function on  M  ⊗R  N ; show that  D   ⊆  ker()From the linearity of  B ,

    B(m + m, n) = B(m, n) + B(m, n), B(m, n + n) = B(m, n) + B(m, n)

    rB(m, n) = B(rm,n) = B(m,rbn) (3.2.10a)

    therefore

    (δ (m+m,n)) = (δ (m,n)) + (δ (m,n)), (δ (m,n+n)) = (δ (m,n)) + (δ (m,n))

    r(δ (m,n)) = (δ (rm,n)) = (δ (m,rn)).   (3.2.11a)

    Since   is linear, these conditions must hold as well:

    (δ (m+m,n)) = (δ (m,n) + δ (m,n)), (δ (m,n+n)) = (δ (m,n) + δ (m,n)),

    (rδ (m,n)) = (δ (rm,n)) = (δ (m,rn)).   (3.2.12a)

    Therefore the kernel of     contains all the generators of the submodule   D, so   will induce a linear map  L :  F R(M  × N )/D  → P   where  L(δ (m,n) + D) =  (δ (m,n)) =B(m, n). Since   F R(M  × N )/D   =   M  ⊗R N   and   δ (m,n) +  D   =   m ⊗ n  there for the

    diagram below commutes as such:

    M  × N M  ⊗R N 

    B   L

  • 8/9/2019 Quantizing Analysis

    23/37

    This diagram shows that every linear map  B  out of  M  × N  comes from a linearmap  L  out of  M  ⊗R N   such that  L(m ⊗ n) =  B(m, n),  ∀m  ∈  M, n  ∈  N . Now weshow 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  F R(M  × N ) tells us

    that every element of  F R(M  × N ) is a finite sum

    r1δ (m1,n1) + r2δ (m1,n2) + · · · + rk2δ (mk,nk).   (3.2.13)

    The reduction map  F R(M  × N ) →  F R(M  × N )/D =  M  ⊗R N   is linear, so everyelement of  M  ⊗R N   is a finite sum of the form

    r1(m1 ⊗ n1) + r2(m1 ⊗ n2) + · · · + rk2(mk ⊗ nk).   (3.2.14)

    Therefore this means the elements of  m⊗n in M ⊗RN  span it as a  R-module. Thelinear maps out of  M  ⊗R N  are determined by the values of all the elements  m ⊗ n,

    thus there is at most one linear map  L  :  M  ⊗R N  → P   such that  m ⊗ n → B(m, n).

    Definition 22.  Tensors in  M ⊗R N  that have the form  m ⊗ n are called  elementary tensors.

    Similar to linear algebra elements of the the tensor product, or tensors, are gen-erally linear combinations of elementary tensors.

    We will now prove two simple theorems that run parallel with our idea of linearalgebra.

    Theorem 3.11.  Le  M   and  N  be  R-modules with respective spanning sets  {xi}i∈I  and 

    {y j} j∈J . The tensor product  M  ⊗R N  is spanned linearly by the elementary tensors xi ⊗ y j.

    Proof.  Since an elementary tensor in  M  ⊗R N  has the form  m ⊗ n, we write  m andn as spanned by the sets:

    m =i

    aixi, n = j

    b jy j   (3.2.15)

    where ai =  b j  = 0 for all but finitely many  i  and  j . Since ⊗   is bilinear we have

    m ⊗ n = i aixi ⊗  j b jy j  = i,j aib jxi ⊗ y j   (3.2.16)which is a linear combination of the tensors  xi ⊗ y j. Since every tensor is a sum

    of elementary tensors, the  xi ⊗ y j ’s span  M  ⊗R N  as an  R-module.

    Theorem 3.12.   In  M  ⊗R N ,  m ⊗ 0 = 0 × n = 0.

  • 8/9/2019 Quantizing Analysis

    24/37

    Proof.   Since  m ⊗ n is additive in n with m  fixed, we have

    m ⊗ 0 = m ⊗ (0 + 0) = m ⊗ 0 + m ⊗ 0 ⇒ m ⊗ 0 = 0.   (3.2.17)

    The same argument is applied to 0 ⊗ n.

    3.3 Tensor products and the Hilbert-Schmidt Class

    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, TheHilbert space  H   is characterized, up to isomorphism, by the existence of a bilinearmapping 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 intuitiveaspects of it. When  x1 ∈ H1, x2  ∈ H2, we want to view the element  p(x1, x2) ∈ H asa ”product”  x1 ⊗ x2. The linear combinations of such products form an everywhere-dense subspace of  H. The bilinearity of  p  implies that these products satisfy certainlinear 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 thebilinearity of  p. The inner product on  H  satisfies the conditions:

    x1 ⊗ x2, y1 ⊗ y2 = x1, y1x2, y2   (3.3.2a)

    x1 ⊗ x22 = x1 ⊗ x2, x1 ⊗ x2

    = x1, x1x2, x2 = x12x2

    2

    x1 ⊗ x2 = x1x2.   (3.3.2b)

    Our construction of the Hilbert space  H, the elements of  H  are complex-valuedfunctions defined on the product  H1 × H2  and conjugate-linear in both variables. If v1  ∈ H1, v2  ∈ H2,  v1 ⊗ v2   is the function that assigns the value  v1, x1v2, x2  to theelement (x1, x2) ∈ H1 × H2.

    Now, suppose that   H1,..., Hn  are Hilbert spaces and   φ   is a mapping from thecartoon product of all these Hilbert spaces into the scalar field   C.   φ   is called a

  • 8/9/2019 Quantizing Analysis

    25/37

    bounded multilinear functional on  H1 × · · · × Hn   if  φ  is linear in each of its variables,assuming that the other variables remain fixed, and there is a real number c  such that

    |φ(x1,...,xn)| ≤ cx1 · · · xn, x1 ∈ H1,...,xn ∈ Hn.   (3.3.3)

    If this is so then the lest such constant  c   is denoted by the norm of the mappingφ. Then φ  is a continuous mapping from H1 × · · · × Hn → C relative to the producttor the norm topologies on the Hilbert spaces.

    Theorem 3.13.  Suppose that  H1,..., Hn  are Hilbert spaces and  φ  is a bounded mul-tilinear functional on  H1 × · · · × Hn.

    (i) The sum 

    y1∈Y 1

    · · ·yn∈Y n

    |φ(y1,...,yn)|2 (3.3.4)

    has the same finite or infinite value for all orthonormal bases  Y 1   of  H1,...,  Y n   of Hn.

    (ii) If   K1,..., Kn   are Hilbert spaces,   Am   ∈ B (Hm, Km),   (m   = 1,...,n),   ψ   is a bounded multilinear functional on  K1 × · · · Kn, and 

    φ(x1,...,xn) = ψ(A1x1,...,Anxn)

    x1 ∈ H1,...,xn ∈ Hn   (3.3.5)

    then 

    y1∈Y 1

    · · ·yn∈Y n

    |φ(y1,...,yn)|2 ≤ A1

    2 · · · An2z1∈Z 1

    · · ·zn∈Z n

    |φ(z 1,...,z n)|2,   (3.3.6)

    when  Y m   and  Z m  are orthonormal bases of  Hm   and  Km, respectively where  m  =1,...,n.

    Proof.  To prove (i) it is sufficient to show that

    y1∈Y 1

    · · ·yn∈Y n

    |φ(y1,...,yn)|2 ≤

    z1∈Z 1

    · · ·zn∈Z n

    |φ(z 1,...,z n)|2,   (3.3.7)

    , whenever   Y m, Z m   are orthonormal bases of   Hm, m   = 1,...,n. Note that sinceY m, Z m  are orthonormal bases we can represent the basis vectors in one in terms of alinear combination of the other.

    y1 =z1∈Z 1

    y1, z 1z 1,...,yn =

    zn∈Z n

    yn, z nz n   (3.3.8)

    Therefore,

  • 8/9/2019 Quantizing Analysis

    26/37

    y1∈Y 1

    · · ·yn∈Y n

    |φ(y1,...,yn)|2 =

    y1∈Y 1

    · · ·yn∈Y n

    |φ(z1∈Z 1

    y1, z 1z 1,...,

    zn∈Z n

    yn, z nz n)|2

    = y1∈Y 1

    · · · yn∈Y n

    | z1∈Z 1

    · · · zn∈Z n

    y1, z 1 · · · yn, z nφ(z 1,...,z n)|2

    ≤z1∈Z 1

    · · ·zn∈Z n

    z 12 · · · z n

    2|φ(z 1,...,z n)|2

    =z1∈Z 1

    · · ·zn∈Z n

    |φ(z 1,...,z n)|2

    (3.3.9a)

    The third line is arrived by noticing that we can rewrite the  z m   ∈   Z m  basis interms of  ym ∈ Y m

    z 1 =y1∈Y 1

    z 1, y1y1,...,z n =yn∈Y n

    z n, ynyn.   (3.3.10)

    Also note that by definition  ym, z m = z m, ym  and that

    z m2 =

    ym∈Y m

    z m, ymym,

    ym∈Y m

    z m, ymym

    =

    ym∈Y m

    ym, z mym, ym, z mym

    = ym∈Y m

    |ym, z m|2ym2

    =

    ym∈Y m

    |ym, z m|2 (3.3.11a)

    Using the same argument and exchanging the orthonormal bases we can showequality.

    Then for the proof of (ii), we suppose that 1  ≤  m  ≤  n  we choose and fix vectorsy1 ∈ Y 1,...,ym−1 ∈ Y m−1, z m+1 ∈ Z m+1,...,z n ∈ Z m. The mapping

    z  → ψ(A1y1,...,Am−1ym−1, z , z  m+1,...,z n) : Km → C   (3.3.12)is a bounded linear functional on  Km, so ∃w ∈ Km  such that

    ψ(A1y1,...,Am−1ym−1, z , z  m+1,...,z n) = z, w, z  ∈ Km   (3.3.13)

    From Parseval’s equation

  • 8/9/2019 Quantizing Analysis

    27/37

    ym∈Y M 

    |ψ(A1y1,...,Am−1ym−1, z , z  m+1,...,z n)|2

    = ym∈Y M 

    |Amym, w|2 = ym∈Y M 

    |ym, A∗mw|2

    = A∗mw2 ≤ Am

    2w2 = Am2

    zm∈Z m

    |z m, w|2

    = Am2

    zm∈Z m

    |ψ(A1y1,...,Am−1ym−1, z , z  m+1,...,z n)|2.   (3.3.14a)

    A further summation now yields

    y1∈Y 1

    · · · ym∈Y m

    zm+1∈Z m+1

    · · · zn∈Z n

    |ψ(A1y1,...,A

    mym

    , z m+1

    ,...,z n

    )|2

    ≤ Am2y1∈Y 1

    · · ·

    ym−1∈Y m−1

    zm∈Z m

    · · ·

    zn∈Z n

    |ψ(A1y1,...,Am−1ym−1, z m,...,z n)|2

    (3.3.15)

    Therefore

    y1∈Y 1

    · · ·yn∈Y n

    |φ(y1,...,yn)|2 =

    y1∈Y 1

    · · ·yn∈Y n

    |ψ(A1y1,...,Anyn)|2

    ≤ An2 y1∈Y 1

    · · · yn−1∈Y n−1

    zn∈Z n

    |ψ(A1y1,...,An−1yn−1, z n)|2

    ≤ An−12An

    2

    ×y1∈Y 1

    · · ·

    yn−2∈Y n−2

    zn−1∈Z n−1

    zn∈Z n

    |ψ(A1y1,...,An−2yn−2, z n−1, z n)|2

    ≤ · · · ≤ A12 · · · An

    2z1∈Z 1

    · · ·zn∈Z n

    |ψ(z 1,...,z n)|2 (3.3.16)

    With  H1,..., Hn   Hilbert spaces, a mapping  φ   :  H1 × · · · × Hn   →  C   is called aHilbert-Schmidt functional on H1 × · · · × Hn   if it si a bounded multilinear functional,and the sum of Eq(3.3.4) is finite for any choice of the orthonormal bases  Y 1   in  H1,..., Y n   in Hn.

    Theorem 3.14.   If  H1,..., Hn  are Hilbert spaces, the set  HSF   of all Hilbert-Schmidt  functionals on  H1 × · · · × Hn  is itself a Hilbert space when the linear structure, inner product, and norm are defined by 

  • 8/9/2019 Quantizing Analysis

    28/37

    (aφ + bψ)(x1,...,xn) = aφ(x1,...,xn) + bψ(x1,...,xn) (3.3.17)

    φ, ψ = y1∈Y 1 · · · yn∈Y n φ(y1,...,yn)ψ(y1,...,yn) (3.3.18)φ2 =

     y1∈Y 1

    · · ·yn∈Y n

    |φ(y1,...,yn)|2

    12

    ,   (3.3.19)

    respectively, where   Y m   is an orthonormal basis in   Hm,   m   = 1,...,n. The sum given by Eq(3.3.18) is absolutely convergent, and the inner product and norm do not depend on the choice of the orthonormal bases  Y 1,...,Y n.

    For each  v(1)   in  H1, ...   v(n)  in  Hn  the equation 

    φv(1),...,v(n)(x1,...,xn) = x1, v(1) · · · xn, v(n), (x1 ∈ H1,...,xn ∈ Hn) (3.3.20)

    defines an element  φv(1),...,v(n)   of  HSF , and 

    φv(1),...,v(n), φw(1),...,w(n) = w(1), v(1) · · · w(n), v(n)   (3.3.21)

    φv(1),...,v(n)2  = v(1) · · · v(n).   (3.3.22)

    The set  {φy(1),...,y(n)   :  y(1)  ∈  Y 1,...,y(n)  ∈  Y n}   is an orthonormal basis of  HSF .There is a unitary transformation  U   from  HSF   onto  l2(Y 1 × · · · × Y n), such that  U φis the restriction  φ|Y 1 × · · · × Y n  when  φ ∈ HS F  .

    Proof.  Choose an orthonormal basis  Y m  in  Hm, m = 1,...,n, and then associate witheach bounded multilinear functional φ  on  H1 × · · · × Hn  the complex-valued functionUφ   obtained by restricting  φ  to  Y 1 × · · · × Y n. Remember that the condition for aHilbert-Schmidt functional is that if  φ  is a Hilbert-Schmidt functional if and only if 

    U φ ∈ l2(Y 1 × · · · × Y n).   (3.3.23)

    If  U φ = 0, then

    φ(y1,...,yn) = 0   y1  ∈ Y 1,...,yn ∈ Y n.   (3.3.24)

    Because  Y m   is an orthonormal basis, its closed linear span is  Hm, then it followsfrom the multi linearity and continuity of  φ  that φ vanishes throughout H1× · · · × Hn.

    Let   φ, ψ   be Hilbert-Schmidt functionals on   H1  × · · · × Hn, then the same istrue of   aφ +  bψ   for all scalars   a, b;   aφ + bψ   is a bounded multilinear functional,Uφ,Uψ ∈ l2(Y 1 × · · · × Y n) and thus

    U (aφ + bψ) = aU φ + bUψ ∈ l2(Y 1 × · · · × Y n) (3.3.25)

  • 8/9/2019 Quantizing Analysis

    29/37

    The sum, given by Eq(3.3.18), can be rewritten in the form

    y∈Y 1×···×Y n

    (U φ)(y)(U ψ)(y) (3.3.26)

    and it is absolutely convergent with the sum   Uφ,Uψ, the inner product inl2(Y 1 × · · · × Y n) of the two restricted functionals  U φ  and  Uψ.

    The set HSF  of all Hilbert-Schmidt functionals on H1×···×Hnis a complex vectorspace. Eq(3.3.18) then defines an inner product on HS F  the restriction  U |HSF   is aone-to-one linear mapping from HSF   into  l2(Y 1 × · · · × Y n), and Uφ,Uψ =  ∠φ, ψwhen  φ, ψ ∈ HSF . The inner product on  l2(Y 1 × · · · × Y n) is definite, so is that onHSF ; if  φ ∈ HSF   and φ, φ = 0, we have  Uφ,Uφ = 0, whence  U φ = 0 ⇒  φ  = 0.From this we see that HSF   is a pre-Hilbert space, and it is apparent from Eq(3.3.18)that the norm, denoted   2   in HSF  is given by Eq(3.3.19).

    From Theorem 3.13, this norm is independent of the choice of the orthonormal

    bases Y 1,..,Y n; this is the same of the inner product on  HSF .Now we will show that  U   brings  HSF  onto the whole of the   l2   space. Let f   ∈

    l2(Y 1  × · · · × Y n) and   xm   ∈ Hm, m   = 1,...,n, the Cauchy-Schwarz inequality andParseval equation gives

    y1∈Y 1

    · · ·yn∈Y n

    |f (y1,...,yn)x1, y1 · · · xn, yn|

     y1∈Y 1

    · · ·yn∈Y n

    |f (y1,...,yn)|2

    12

    × 

    y1∈Y 1

    · · ·yn∈Y n

    |x1, y1|2 · · · | xn, yn|

    2 12

    = f 

     y1∈Y 1

    |x1, y1|2

    12

    · · ·

     yn∈Y n

    |xn, yn|2

    12

    = f x1 · · · xn   (3.3.27a)

    From this, the equation

    φ(x1,...,xn) = y1∈Y 1 · · · yn∈Y n f (y1,...,yn)x1, y1 · · · xn, yn   (3.3.28)defines a bounded multilinear functional  φ  on  H1 × · · · × Hn, with  φ ≤ f .

    The orthonormality of the sets  Y 1,...,Y n, leads to the fact that

    (Uφ)(y1,...,yn) = φ(y1,...,yn) = f (y1,...,yn)   y1 ∈ Y 1,...,yn ∈ Y n   (3.3.29)

  • 8/9/2019 Quantizing Analysis

    30/37

    so U φ =  f . Furthermore φ ∈ HS F   since U φ ∈ l2(Y 1 × · · · × Y n), whence U  carriesHSF  onto the  l2  space.

    Since  U   is a norm preserving linear mapping from  HSF   onto   l2(Y 1 × · · · × Y n)completeness of the l2 space entails completeness of HSF ; so HSF  is a HIlbert space,

    and U   is a unitary operator.When v(1) ∈ H1,...,v(n) ∈ Hn,  φv(1),...,v(n)  is a multilinear functional on  H1 × · ·

    · × Hn, and is bounded since

    |φv(1),...,v(n)(x1,...,xn)| ≤ v(1) · · · v(n)x1 · · · xn   (3.3.30)

    by the Cauchy-Schwarz inequality. Furthermore, Parseval’s equation gives

    y1∈Y 1

    · · ·yn∈Y n

    |φv(1),...,v(n)(y1,...,yn)|2

    = y1∈Y 1

    · · · yn∈Y n

    |y1, v(1)|2 · · · | yn, v(n)|2

    =

     y1∈Y 1

    |y1, v(1)|2

    · · ·

     yn∈Y n

    |yn, v(n)|2

    = v(1)2 · · · v(n)2.   (3.3.31)

    Hence φv(1),...,v(n)  ∈ HS F    and φv(1),...,v(n)2  = v(1) · · · v(n). Using Parseval’sequation again and by absolute convergence,

    φv(1),...,v(n), φw(1),...,w(n)

    =y1∈Y 1

    · · ·yn∈Y n

    φv(1),...,v(n)(y1,...,y(n))φw(1),...,w(n)(y1,...,yn)

    =y1∈Y 1

    · · ·yn∈Y n

    y1, v(1) · · · yn, v(n)w(1), y1 · · · w(n), yn

    =

     y1∈Y 1

    w(1), y1y1, v(1)

    · · ·

     yn∈Y n

    w(n), ynyn, v(n)

    = w(1), v(1) · · · w(n), v(n)   (3.3.32)

    When   y(1)   ∈   Y 1,...,y(n)   ∈   Y n, the orthonormality of   Y 1,...,Y n   implies thatUφy(1),...,y(n)   is the function that takes the value 1 at (y(1),...,y(n)) and 0 elsewhereon Y 1 × · · · × Y n. Therefore

    {U φy(1),...,y(n) :  y(1) ∈ Y 1,...,y(n) ∈ Y n}   (3.3.33)

    is an orthonormal basis of  l2(Y 1 × · · · × Y n), and therefore

  • 8/9/2019 Quantizing Analysis

    31/37

    {φy(1),...,y(n) :  y(1) ∈ Y 1,...,y(n) ∈ Y n}   (3.3.34)

    is a basis of  HSF .

    Let us introduce the notion of the ”conjugate” of a Hilbert space  H. For a ”nor-mal” Hilbert space,  H  has the algebraic structure and inner product defined by themappings

    (x, y) → x + y  : H × H → H   (3.3.35a)

    (a, x) → ax : C × H → H   (3.3.35b)

    (x, y) → x, y : H × H → C.   (3.3.35c)

    However we define the ”conjugate” Hilbert space with a slight twist.

    Definition 23.  The  conjugate Hilbert space  H̄   is the same set  H, with the alge-braic structure and inner product defined by the mappings:

    (x, y) → x + y : H × H → H   (3.3.36a)

    (a, x) → ā·x : C × H → H   (3.3.36b)

    (x, y) → x, y− : H × H → C   (3.3.36c)

    where 

    ā·x = āx   and  x, y− = y, x.   (3.3.37)

    It is also very easy to see that the conjugate Hilbert space of  H̄  is  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 ithas the same property relative to the conjugate Hilbert space. If   H1   and   H2   areHilbert spaces and  T   is a mapping from the set  H1   into the set  H2, the linearity of T   : H1 → H2 is equivalent to linearity of  T   :  H̄1 →  H̄2, and corresponds to conjugate-linearity of  T   : H1 →  H̄2  and of  T   :  H̄1 → H2. Of course, continuity of  T   is the samein all four situations, when  T   is linear the operators have the same bound, since thenorm on H j   is the same as that on  H̄ j.

    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 

    L(x1,...,xn) ≤ cx1 · · · xn.   (3.3.38)

    The least such constant  c  is denoted by  L.

  • 8/9/2019 Quantizing Analysis

    32/37

    By a weak Hilbert-Schmidt mapping from   H1  × · · · × Hn   into   K, we mean abounded multilinear mapping  L  with the properties:

    (i)

    Lu(x1,...,xn) = L(x1,...,xn), u ∀u ∈ K   (3.3.39)

    is a Hilbert-Schmidt functional on  H1 × · · · × Hn.(ii) There is a real number  d  such that Lu2 ≤ du  for each  u ∈ K.When these conditions are fulfilled, the least possible value of the constant  d   is

    denoted by L2.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,  T  = L2.

    (ii) If  H and  p have the properties attributed in (i) to  H  and  p, there is a unitary transformation  U   from  H  onto  H such that  p = U p.

    (iii) If  vm, wm ∈ Hm  and  Y m  is an orthonormal basis of  Hm, m = 1,...,n, then 

     p(v1,...,vn), p(w1,...,wn) = v1, w1 · · · vn, wn,   (3.3.40)

    the set   { p(y1,...,yn) :   y1   ∈   Y 1,...,yn   ∈   Y n}   is an orthonormal basis of   H, and  p2  = 1.

    Proof.   Let  H̄m   be the conjugate Hilbert space of   Hm   and let   H   be the set of allHilbert-Schmidt functionals on  H̄1 × · · · ×  H̄n  with the Hilbert space structure givenin Theorem 3.14. When  v(1) ∈ H1,...,v(n) ∈ Hn, let  p(v(1),...,v(n)) be the Hilbert-Schmidt functional   φv(1),...,v(n)   defined on the cartesian product of   H̄1  × · · · ×  H̄nby

    φv(1),...,v(n)(x1,...,xn) = x1, v(1)− · · · xn, v(n)

    = v(1), x1 · · · v(n), xn   (3.3.41)

    Because we let  Y  j  be an orthonormal basis of  H j, j  = 1,...,n Theorem 3.14 then

    says that the set  { p(y1,...,yn) :  y1   ∈  Y 1,...,yn   ∈  Y n}   is an orthonormal basis of  H,and that

     p(v1,...,vn), p(w1,...,wn) = w1, v1− · · · wn, vn

    = v1, w1 · · · vn, wn,   (3.3.42a)

     p(v1,...,vn)2 = v1 · · · vn.   (3.3.42b)

  • 8/9/2019 Quantizing Analysis

    33/37

    From above we have  p  :  H1 × · · · × Hn  → H  is a bounded multilinear mapping:we shall prove next that it is a weak Hilbert-Schmidt mapping. Suppose that φ ∈ H,and consider the bounded multilinear functional  pφ : H1 × · · · × Hn → C defined by

     pφ(x1,...,xn) =  p(x1,...,xn), φ.   (3.3.43)

    With y(1) ∈ Y 1,...,y(n) ∈ Y n, orthonormality of the bases implies that  φy(1),...,y(n)takes the value 1 at (y(1),...,y(n)) and 0 elsewhere on  Y 1 × · · · × Y n. Thus

     pφ(y(1),...,y(n)) =  p(y(1),...,y(n)), φ = φy(1),...,y(n), φ

    =y1∈Y 1

    · · ·yn∈Y n

    φy(1),...,y(n)(y1,...,yn)φ(y1,...,yn)

    = φ(y(1),...,y(n)),   (3.3.44a)

    = y(1)∈Y 1

    · · · y(n)∈Y n

    | pφ(y(1),...,y(n))|2 = φ22.   (3.3.44b)

    From this we have proven that  pφ is a Hilbert-Schmidt functional on H1 × · · · × Hnand that  pφ2 =  phi2; so p  : H1×···×Hn → H is a weak Hilbert-Schmidt mappingwith  p2 = 1.

    Next, suppose that L  is a weak Hilbert-Schmidt mapping from H1 × · · · × Hn  intoanother Hilbert space  K. Let  u ∈ K  and  Lu   is the Hilbert-Schmidt functional givenin Definition 23, while  φ  ∈ H  and let  F  be a finite subset of  Y 1 × · · · × Y n  then wehave using Cauchy-Schwarz

    |

    (y1,...,yn)∈F

    φ(y1,...,yn)L(y1,...,yn), u|

    (y1,...,yn)∈F

    |φ(y1,...,yn)||Lu(y1,...,yn)|

      (y1,...,yn)∈F

    |φ(y1,...,yn)|2

    12   (y1,...,yn)∈F

    |Lu(y1,...,yn)|2

    12

    ≤ Lu2

      (y1,...,yn)∈F

    |φ(y1,...,yn)|2

    12

    ≤ uL2

      (y1,...,yn)∈F

    |φ(y1,...,yn)|2

    12

    (3.3.45)

    Hence

  • 8/9/2019 Quantizing Analysis

    34/37

    (y1,...,yn)∈F

    φ(y1,...,yn)L(y1,...,yn)

    ≤ L2

      (y1,...,yn)∈F

    |φ(y1,...,yn)|2 12

    (3.3.46)

    Since

    y1∈Y 1

    · · ·yn∈Y n

    |φ(y1,...,yn)|2 = φ22 < ∞   (3.3.47)

    it then follows from Eq(3.3.46) and the Cauchy criterion that the, unordered, sum

    y1∈Y 1

    · · · yn∈Y n

    φ(y1,...,yn)L(y1,...,yn) (3.3.48)

    converges to an element  T φ ∈ K, and T φ ≤ L2 phi2. Thus  T   is a boundedlinear operator from  H   into  K, and  T  ≤ L2. When  y(1) ∈  Y 1,...,y(n) ∈  Y n, wehave

    T p(y1,...,yn) = T φy(1),...,y(n)

    =y1∈Y 1

    · · ·yn∈Y n

    φy(1),..,y(n)(y1,...,yn)L(y1,...,yn)

    = L(y(1),...,y(n)).   (3.3.49)

    Both L and T p are bounded and multiline and Y m 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(Y 1 × · · · × Y n) of  H.   ∀u ∈ K,Parseval’s equation gives

    Lu22  =

    y1∈Y 1

    · · ·yn∈Y n

    |L(y1,...,yn), u|2

    = y1∈Y 1

    · · · yn∈Y n

    |T p(y1,...,yn), u|2

    =y1∈Y 1

    · · ·yn∈Y n

    | p(y1,...,yn), T ∗u|2

    = T ∗u2 ≤ T 2u2; (3.3.50)

    so we have L2 ≤ T , and thus  L2 = T .

  • 8/9/2019 Quantizing Analysis

    35/37

    We now prove (ii) of the theorem. Suppose that  H and p : H1 × · · · × Hn → H

    have the properties given in (i). When we have that K = H and L  =  p, the equationL  =  T p is satisfied when  T   is the identity operator on  H, and also when  T   is theprojection from  H onto the closed subspace [ p(H1 × · · · × Hn)] generated by the

    range p(H1 × · · · × Hn) of  p. From the uniqueness of  T ,

    [ p(H1 × · · · × Hn)] = H (3.3.51)

    furthermore,

     p2 = L2  = T  = I  = 1.   (3.3.52)

    Using a similar argument and letting   K   =   H and   L   =   p, it follows from theproperties of the Hilbert space   H   and   p  given in (i), that theres a hounded linearoperator  U   : H → H” such that  p = U p  and

    u = L2 =  p2 = 1 (3.3.53)

    The roles of H, p and H, p can be reversed in this argument, so there is a boundedlinear operator  U  from H into  H  such that  p =  U  p and U  = 1. Since

    U U p(x1,...,xn) = U  p(x1,...,xn) = p(x1,...,xn),   ∀x1 ∈ H1,...,xn ∈ Hn,   (3.3.54)

    while

    [ p(H1 × · · · × Hn)] = H   (3.3.55)

    it follows that U U  is the identity operator on H; and similarly U U  is the identityoperator on H. Finally,

    x = U U x ≤ U x ≤ x,   ∀x ∈ H   (3.3.56)

    so U x = x, and U  is an isomorphism from  H onto H.

    4 Stinespring’s Dilation Theorem

    Theorem 4.1.   Let  A  be a unital  C ∗- algebra, and let  φ : A → B(H)  be a completely positive map. Then there exists a Hilbert space  K, a unital  ∗-homomorphism  π  : A →B(K), and a bounded operator  V   : H → K  with  φ(1) = V 2 such that 

    φ(a) = V ∗π(a)V    (4.0.57)

  • 8/9/2019 Quantizing Analysis

    36/37

    Proof.   Consider the algebraic tensor product  A ⊗ H, of the unital   C ∗-algebra andthe Hilbert Space  H, and define a symmetric bilinear function  ,  on this space bysetting

    a ⊗ x, b ⊗ y = φ(b∗a)x, yH,   (4.0.58)

    and extending linearly, where , H is the inner product on H. From the definition,φ  is completely positive thereby ensuring that ,   is positive semidefinite since

      n j=1

    a j ⊗ x j,n

    i=1

    ai ⊗ xi

    =

    φn((a

    ∗a))

    x1...

    xn

    ,

    x1...

    xn

    H(n)

    ≥ 0,   (4.0.59)

    where , H(n)  denotes the inner product not the direct sum  H(n) of  n  copies of  H,

    given by

    x1...

    xn

    ,

    y1...

    yn

    H(n)

    = x1, y1H + · · · + xn, ynH   (4.0.60)

    A result of positive semidefinite bilinear forms is that they satisfy the Cauchy-Schwarz inequality,

    |u, v|2 ≤ u, u · v, v   (4.0.61)

    Therefore we have that

    {u ∈ A ⊗ H|u, u = 0} = {u ∈ A ⊗ H|u, v = 0∀v ∈ A ⊗ H}   (4.0.62)

    is a subspace,   N , of  A ⊗ H. The induced bilinear form on the quotient spaceA ⊗ H/ N   defined by

    u + N , v + N = u, v   (4.0.63)

    will be an inner product. We let K denote the Hilbert space that is the completionof the inner product space  A ⊗ H/ N .

    If  a ∈ A, define a linear map  π(a) : A ⊗ H → A ⊗ H  by

    π(a)

    ai ⊗ xi

    =

    (aai) ⊗ xi   (4.0.64)

    A matrix factorization shows that the following inequality in  M n(A)+ is satisfied:

    (a∗i a∗aa j) ≤ a

    ∗a · (a∗i a j),   (4.0.65)

  • 8/9/2019 Quantizing Analysis

    37/37

    and consequently,

    π(a) a j ⊗ x j, π(a) ai ⊗ xi=i,j

    φ(a∗i a∗aa j)x j, xiH ≤ a

    ∗a ·i,j

    φ(a∗i a j)x j, xiH

    = a2 ·

    a j ⊗ x j ,

    ai ⊗ xi

    .   (0.0.12)

    Therefore, π(a) leave N   invariant and consequently induces a quotient linear trans-formation on A ⊗ H/ N , which we will still denote by π(a). The above inequality alsoshows that  π(a) is bounded with  π(a) ≤ a. Thus,  π(a) extends to a boundedlinear operator on K, which we will still denote by π(a). It is straightforward to verify

    that the map  π  : A → B(K) is a unital  ∗   - homomorphism.Now define  V   : H → K  via

    V (x) = 1 ⊗ x + N    (4.0.67)

    Then V   is bounded, since

    V x2 = 1 ⊗ x, 1 ⊗ x = φ(1)x, xH ≤  phi(1) · x2.   (4.0.68)

    Indeed, it is clear that V 2 = sup{φ(1)x, xH  : x ≤ 1} = φ(1).To complete the proof, we only need to observe that

    V ∗π(a)V x , yH = π(a)1 ⊗ x, 1 ⊗ yK = φ(a)x, yH,   ∀x, y ∈ H,   (4.0.69)

    and so  V ∗π(a)V   = φ(a).