Next Article in Journal
A Stroll through the Loop-Tree Duality
Next Article in Special Issue
Modeling Neuronal Systems as an Open Quantum System
Previous Article in Journal
The COVID-19 Vaccine Preference for Youngsters Using PROMETHEE-II in the IFSS Environment
Previous Article in Special Issue
The Brain and the New Foundations of Mathematics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Parts and Composites of Quantum Systems

by
Stanley P. Gudder
Department of Mathematics, University of Denver, Denver, CO 80208, USA
Symmetry 2021, 13(6), 1031; https://doi.org/10.3390/sym13061031
Submission received: 7 May 2021 / Revised: 28 May 2021 / Accepted: 1 June 2021 / Published: 8 June 2021
(This article belongs to the Special Issue Quantum Information Applied in Neuroscience)

Abstract

:
We consider three types of entities for quantum measurements. In order of generality, these types are observables, instruments and measurement models. If α and β are entities, we define what it means for α to be a part of β . This relationship is essentially equivalent to α being a function of β and in this case β can be employed to measure α . We then use the concept to define the coexistence of entities and study its properties. A crucial role is played by a map α ^ which takes an entity of a certain type to one of a lower type. For example, if I is an instrument, then I ^ is the unique observable measured by I . Composite systems are discussed next. These are constructed by taking the tensor product of the Hilbert spaces of the systems being combined. Composites of the three types of measurements and their parts are studied. Reductions in types to their local components are discussed. We also consider sequential products of measurements. Specific examples of Lüders, Kraus and trivial instruments are used to illustrate various concepts. We only consider finite-dimensional systems in this article. Finally, we mention the role of symmetry representations for groups using quantum channels.

1. Introduction

Two important operations on quantum systems are the formations of parts and composites. In a rough sense, these operations are opposites to each other. The parts of a measurement α are smaller components of α in the sense that they can be simultaneously measured by α . A composite system is a combination of two or more other systems. This combination is formed using the tensor product H = H 1 H 2 , where H 1 and H 2 are the Hilbert spaces describing two subsystems. The composite system contains more information than the individual systems because H describes how H 1 and H 2 interact. We can reduce measurements on H to simpler ones on H 1 and H 2 but information is lost in the process.
Section 2 presents the basic definitions that are needed in the sequel. Three types of quantum measurements are considered. In order of generality, these types are observables, instruments and measurement models. At the basic level is an observable A which is a measurement whose outcome probabilities tr ( ρ A x ) are determined by the state ρ of the system. At the next level is an instrument I . We think of I as an apparatus that can be employed to measure an observable I ^ . Although I ^ is unique, there are many instruments that can be used to measure an observable. Moreover, I gives more information than I ^ because, depending on the outcome x, I updates the input state ρ to the output state I x ( ρ ) / tr ( ρ I ^ x ) . At the highest level is a measurement model M that measures a unique instrument M ^ . Again, there are many measurement models that measure an instrument and M contains more detailed information. For conciseness, we call these types of measurements entities. We should mention that all the quantum systems in this article are assumed to be finite-dimensional.
Section 3 considers system parts. If α and β are entities, we define what it means for α to be a part of β and when this is the case, we write α β . If α β and β α , we say that α and β are equivalent. We show that α β implies α ^ β ^ and that → is a partial order to within equivalence. The relation α β is the same as α being a function of β or β ^ and in this case, β can be employed to measure α . We then use this concept to define the coexistence of entities and study its properties. We show that joint measurability is equivalent to coexistence. We then introduce the sequential products of observables and use this concept to illustrate parts of entities.
Section 4 discusses composite systems. These are constructed by taking the tensor product H = H 1 H 2 , where H 1 , H 2 are the Hilbert spaces of the systems being combined. Composites of the three types of measurements and parts of these composites are studied. Reductions of types into their local components are discussed. Specific examples of Lüders, Kraus and trivial instruments are employed to illustrate various concepts.

2. Basic Definitions

This section discusses the basic concepts and definitions that are needed in the sequel. Since these ideas are well developed in the literature [1,2,3,4,5], we shall proceed quickly and leave details and motivation to the reader’s discretion. In this article, we shall only consider finite-dimensional complex Hilbert spaces H. Let L ( H ) be the set of linear operators on H. For S , T L ( H ) , we write S T if ϕ , S ϕ ϕ , T ϕ for all ϕ H . We define the set of effects by
E ( H ) = a L ( H ) : 0 a 1
where 0 , 1 are the zero and identity operators, respectively. Effects correspond to yes–no measurements and when the result of the measurement a is yes, we say that a occurs. The complement of a E ( H ) is a = 1 a and a occurs if and only if a does not occur. A one-dimensional projection P ϕ = ϕ ϕ , where ϕ = 1 is an effect called an atom. We call ρ E ( H ) a partial state if tr ( ρ ) 1 and ρ is a state if tr ( ρ ) = 1 . We denote the set of partial states by S p ( H ) and the set of states by S ( H ) . If ρ S ( H ) , a E ( H ) , we call P ρ ( a ) = tr ( ρ a ) the probability that a occurs in the state ρ [1,2,3]. For a , b E ( H ) , their sequential product is the effect a b = a 1 / 2 b a 1 / 2 , where a 1 / 2 is the unique square root of a [6,7,8]. We interpret a b as the effect that results from first measuring a and then measuring b. We also call a b the effect b conditioned on the effect a and write ( b a ) = a b .
Let Ω A be a finite set. A (finite) observable with outcome-space Ω A is a subset:
A = A x : x Ω A E ( H )
satisfying x Ω A A x = 1 . We denote the set of observables on H by O ( H ) . If B = B y : y Ω B is another observable, we define the sequential product A B O ( H ) [8,9,10] to be the observable with outcome-space Ω A × Ω B given by
A B = A x B y : ( x , y ) Ω A × Ω B
We also define the observable B conditioned by A as
( B A ) = ( B A ) y : y Ω B E ( H )
where ( B A ) y = x Ω A ( A x B y ) . If A O ( H ) , we define the effect-valued measure (or POVM) X A X from 2 Ω A to E ( H ) by A X = x X A x and we also call X A X an observable [2,3,8]. Moreover, we have the observables:
( A B ) Δ = ( x , y ) Δ ( A x B y )
and:
( B A ) Y = x Ω A ( A x B Y )
If ρ S ( H ) and A O ( H ) , the probability that A has an outcome in X Ω A when the system is in state ρ is P ρ ( A X ) = tr ( ρ A X ) . Notice that X P ρ ( A X ) is a probability measure on Ω A . We call:
P ρ ( A X then B Y ) = tr ρ ( A B ) X × Y
the joint probability of A X then B Y [8,9,10].
An operation is a completely positive map A : S p ( H ) S p ( H ) [1,2,3]. Any operation has a Kraus decomposition:
A ( ρ ) = i = 1 n S i ρ S i *
where S i L ( H ) with i = 1 n S i * S i 1 . An operation A is a channel if A ( ρ ) S ( H ) for all ρ S ( H ) . In this case, i = 1 n S i * S i = 1 and we denote the set of channels on H by C ( H ) . Notice that if a E ( H ) , then ρ ( ρ a ) = a ρ is an operation and if A O ( H ) , then ρ ( ρ A ) = x Ω A ( A x ρ ) is a channel. The simplest example of a channel has the form A U ( ρ ) = U ρ U * where U is a unitary operator on H. Letting U ( H ) be the group of unitary operators on H, we have that A I = I and A U 1 U 2 = A U 1 A U 2 for all U 1 , U 2 U ( H ) . In particular, if we have a symmetry group G U ( H ) for the system, then U A U , U G , gives a symmetry representation of G. For a finite set Ω I , a (finite) instrument with outcome-space Ω I is a set of operations I = I x : x Ω I satisfying C I = x Ω I I x C ( H ) [1,2,3,11]. Defining I X = x X I x for X Ω I , we see that X I X is an operation-valued measure on H that we also call an instrument. We denote the set of instruments on H by In ( H ) . We say that I In ( H ) measures A O ( H ) if Ω A = Ω I and:
P ρ ( A X ) = tr I X ( ρ )
for every ρ S ( H ) , X Ω A . There is a unique A O ( H ) that I measures and we write A = I ^ [1,2,11]. For I , J In ( H ) , we define the product instrument with outcome space Ω I × Ω J by
( I J ) ( x , y ) ( ρ ) = J y I x ( ρ )
for every ρ S ( H ) . We also define the conditioned instrument with outcome-space Ω J by
( J I ) y = x Ω I ( I J ) ( x , y ) = J y C I ( ρ )
We conclude that:
( I J ) Δ ( ρ ) = ( x , y ) Δ J y I x ( ρ )
for all Δ Ω I × Ω J and:
( J I ) Y ( ρ ) = y Y J y C I ( ρ )
for all Y Ω J [8,9,10].
A finite measurement model (MM) is a 5-tuple M = ( H , K , η , ν , F ) where H, K are finite-dimensional Hilbert spaces called the base and probe systems, respectively, η S ( K ) is an initial probe state, ν C ( H K ) is a channel describing the measurement interaction between the base and probe systems, and F O ( K ) is the probe (or meter) observable [1,2,12]. We say that M measures the model instrument M ^ In ( H ) where M ^ is the unique instrument satisfying:
M ^ X ( ρ ) = tr K ν ( ρ η ) ( I F X )
for all ρ S ( H ) , X Ω F . In (2), tr K is the partial trace over K [2,3]. We also say that M measures the model observable M .
We thus have three levels of abstraction. At the basic level is an observable A which is a measurement whose outcome probabilities tr ( ρ A x ) are determined by the state ρ of the system. At the next level is an instrument I . We think of I as an apparatus that can be employed to measure an observable I ^ . Although I ^ is unique, there are many instruments that can be used to measure an observable. Moreover, I gives more information than I ^ because, depending on the outcome x (or event X), I updates the input state ρ to the output partial state I x ( ρ ) (or I X ( ρ ) ). At the highest level is a measurement model M that measures a unique model instrument M ^ and a unique model observable M . Again, there are many M M s that measure any instrument or observable and M contains more detailed information on how the measurement is performed.

3. System Parts

We begin by discussing parts of systems at the three levels considered in Section 2. We then show how parts can be used to define coexistence at these levels and even between levels. We also show that coexistence is equivalent to simultaneous measurability.
An element at one of the three levels discussed in Section 2 is called an entity. The three levels are said to be the types 1, 2 and 3, respectively. The concept of an entity being part of another entity was originally introduced in [12,13]. If A , B O ( H ) , we say that A is part of B (and write A B ) if there exists a surjection f : Ω B Ω A such that A x = B f 1 ( x ) for all x Ω A . We then write A = f ( B ) . It follows that A X = B f 1 ( X ) for all X Ω A and that:
A X = B y : f ( y ) X
If I , J In ( H ) , we say that I is part of J (and write I J ) if there exists a surjection f : Ω J Ω I such that I x = J f 1 ( x ) for all x Ω J . We then write I = f ( J ) and an equation analogous to (3) holds. For M M s M 1 = ( H , K , η , ν , F 1 ) and M 2 = ( H , K , η , ν , F 2 ) , we say that M 1 is part of M 2 (and write M 1 M 2 ) if F 1 F 2 . It follows that F 1 = f ( F 2 ) and we write M 1 = f ( M 2 ) . We can also define “part of” for entities of different types. An observable A O ( H ) is part of I In ( H ) (written A I ) if A I ^ and A is part of  M (written A M ) if A M ^ which is equivalent to A M . Finally, we say that I is part of M (written I M ) if I M ^ . Two entities α and β are equivalent (written α β ) if α β and β α . It is easy to check that ≅ is an equivalence relation and that α β if and only if α = f ( β ) for f a bijection. Our first result summarizes properties possessed by “part of”. Some of these properties have been verified in [13], however, we give the full proof for completeness.
Theorem 1.
(a) If α , β are of types 2 or 3 and α β , then α ^ β ^ ; (b) f ( I ^ ) = f ( I ) and f ( M ^ ) = f ( M ) ; (c) If α , β , γ are of the same type and α = g ( β ) , β = f ( γ ) , then α = ( g f ) ( γ ) ; (d) The relation → is a partial order to within equivalence; (e) If α and β are of different types and α β , then α = β ^ 1 where β 1 β .
Proof. 
(a) Let I , J In ( H ) with I J . Then, there exists a surjection f : Ω J Ω I such that I = f ( J ) . We now show that I ^ = f ( J ^ ) . Indeed, for any ρ S ( H ) , x Ω I we have that:
tr ( ρ I ^ x ) = tr I x ( ρ ) = tr J f 1 ( x ) ( ρ ) = tr ρ J ^ f 1 ( x ) = tr ρ f ( J ^ ) x
Hence, I ^ = f ( J ^ ) so I ^ J ^ . Let M 1 = ( H , K , η , ν , F 1 ) , M 2 ( H , K , η , ν , F 2 ) be M M s where F 1 = f ( F 2 ) . Then, for any ρ S ( H ) , x Ω F 1 we have that:
M ^ 1 , x ( ρ ) = tr K ν ( ρ η ) ( I F 1 , x ) = tr K ν ( ρ η ) ( I F 2 , f 1 ( x ) ) = M ^ 2 , f 1 ( x ) ( ρ ) = f ( M ^ 2 )
Hence, M ^ 1 = f ( M ^ 2 ) so M ^ 1 M ^ 2 . If I M , then I M ^ . As before, I ^ M so I ^ M ^ .
(b) This was proved in (a). (c) We prove the result for observables A , B , C and the result for instruments and M M s is similar. We have that A x = B g 1 ( x ) and B y = C f 1 ( y ) . Since g : Ω B Ω A and f : Ω C Ω B , we have that g f : Ω C Ω A . Hence:
A x = B g 1 ( x ) = C f 1 ( g 1 ( x ) ) = C ( g f ) 1 ( x )
Hence, A = ( g f ) ( C ) . (d) We only need to prove that if α β and β γ , then α γ . If α , β , γ are of the same type, the α γ follows from (c). Suppose A , B O ( H ) , I In ( H ) and A B , B I . Then, A B I ^ and these are the same type so A I ^ and hence, A I . Suppose A O ( H )   I , J In ( H ) and A I , I J . Then, A I ^ and I J . By (a), we have I ^ J ^ . Since A , I ^ , J ^ have the same type, A J ^ and hence, A J . Suppose that A I and I M . Then, A I ^ and I M ^ . By (a) I ^ M so A I ^ and I ^ M . Since these are the same type, we have that A M so A M . Similar reasoning holds for the cases I J M and I M 1 M 2 .
(e) If A O ( H ) , I In ( H ) and A I , then A I ^ so A = f ( I ^ ) for some surjection f : Ω J ^ Ω A . By (b), we have that f ( I ^ ) = f ( I ) so letting I 1 = f ( I ) we have that A = f ( I ) = I ^ 1 . Hence, I 1 I . If A M , then A M . By (b), A = f ( M ) = f ( M ^ ) . Letting I = f ( M ^ ) , we have that A = I ^ , I M ^ M . If I M , then I M ^ . By (b) I = f ( M ) = f ( M ) . Letting I 1 = f ( M ) , we have that I = I ^ 1 and I 1 M . □
For an entity α , we denote its set of parts by a ˜ = β : β α . We say that a set A of entities coexist if A a ˜ for some entity α . A coexistent set A a ˜ is thought of as being simultaneously measured by α . A related concept is that of joint measurability [14]. We say that observables A i O ( H ) with outcome sets Ω i , i = 1 , 2 , , n are jointly measurable with joint observable B O ( H ) if Ω B = Ω 1 × × Ω n and for all x i Ω i we have:
A x i i = B ( x 1 , , x i , , x n ) : x j Ω j , j i
We interpret A i as being the ith marginal of B as in classical probability theory [8,9,12]. Similar definitions can be made for the joint measurability of instruments and M M s.
Theorem 2.
A set of observables A i O ( H ) , i = 1 , 2 , , n is jointly measurable if and only if the A i coexists.
Proof. 
If A i : i = 1 , 2 , , n are jointly measurable, there exists a joint observable B O ( H ) satisfying (4). Defining f i : Ω B Ω A i by
f i ( x 1 , , x i , , x n ) = x i
for i = 1 , 2 , , n , then by (4), we have that A x i i = B f i 1 ( x i ) for all x i Ω i . Hence, A i = f i ( B ) , i = 1 , 2 , , n , so A i coexist. Conversely, suppose that A i : i = 1 , 2 , , n coexist so there exists an observable C O ( H ) such that A i C ˜ , i = 1 , 2 , , n . We then have surjections f i : Ω C Ω A i such that A i = f i ( C ) , i = 1 , 2 , , n . Define Ω B = Ω 1 × × Ω n , a surjection h : Ω C Ω B by h ( y ) = f 1 ( y ) , , f n ( y ) and let B = h ( C ) . For i = 1 , 2 , , n , we obtain:
A x i i = C f i 1 ( x i ) = C y : f i ( y ) = x i = C y : f 1 ( y ) , , f n ( y ) = x 1 , , x i , , x n , x j Ω j , j i = C y : h ( y ) = ( x 1 , , x i , , x n ) : x j Ω j , j i = C h 1 ( x 1 , , x i , , x n ) : x j Ω j , j i = h ( C ) ( x 1 , , x i , , x n ) : x j Ω j , j i = B ( x 1 , , x i , , x n ) : x j Ω j , j i
Thus, (4) holds so A i are jointly measurable. □
Theorem 2 also holds for instruments and M M s. An important property of coexistent entities is that they have joint probability distributions Φ ρ for all ρ S ( H ) . For example, if A , B O ( H ) coexist, then A = f ( C ) , B = g ( C ) for some C O ( H ) . Then, for any X Ω A , Y Ω B , the joint probability becomes:
Φ ρ ( A X , B Y ) = tr ρ C z : z f 1 ( X ) g 1 ( Y ) = tr ρ C f 1 ( X ) g 1 ( Y )
As another example, if A , B I , then A , B I ^ so A = f ( I ^ ) , B = g ( I ^ ) for surjections f , g . We then obtain:
Φ ρ ( A X , B Y ) = tr ρ I ^ f 1 ( X ) g 1 ( Y ) = tr I f 1 ( X ) g 1 ( Y ) ( ρ )
We can continue this for many coexistent entities. Moreover, the entities do not need to be of the same type. For instance, suppose A , I J where A = f ( J ^ ) and I = g ( J ) . Then, we have that:
Φ ρ ( A X , I Y ) = Φ ρ f ( J ^ ) X , g ( J ) Y = Φ ρ J ^ f 1 ( X ) , J g 1 ( Y ) = tr J f 1 ( X ) g 1 ( Y ) ( ρ )
For A O ( H ) , we define the probability distribution Φ ρ A ( X ) = tr ( ρ A X ) for all X Ω A , ρ S ( H ) . In a similar way, if I In ( H ) , we define Φ ρ I ( X ) = tr I X ( ρ ) and if M is a M M , then Φ ρ M ( X ) = Φ ρ M ^ ( X ) .
Lemma 1.
If α is an entity and f : Ω α Ω is a surjection, then Φ f ( α ) = Φ α f 1 .
Proof. 
We give the proof for A O ( H ) and the proof for other entities is similar. For x Ω A , ρ S ( H ) we obtain:
Φ ρ f ( A ) ( x ) = tr ρ f ( A ) x = tr ρ A f 1 ( x ) = tr ρ A y : f ( y ) = x = tr ( ρ A y ) : f ( y ) = x = Φ ρ A ( y ) : f ( y ) = x = Φ ρ A f 1 ( x ) = Φ ρ A f 1 ( x )
The result now follows. □
We now consider sequential products of observables.
Theorem 3.
If A , B O ( H ) and h : Ω B Ω is a surjection, then A, ( B A ) and A h ( B ) are parts of A B .
Proof. 
Defining f : Ω A × Ω B Ω A by f ( x , y ) = x we have that:
f ( A B ) x = ( A B ) f 1 ( x ) = ( A B ) ( y , z ) : f ( y , z ) = x = z Ω B ( A B ) ( x , z ) = z Ω B A x B z = A x 1 = A x
Thus, A = f ( A B ) so A A B . Defining g : Ω A × Ω B Ω B by g ( x , y ) = y we obtain:
g ( A B ) y = ( A B ) g 1 ( y ) = ( A B ) ( x , z ) : g ( x , z ) = y = x Ω A ( A B ) ( x , y ) = x Ω A A x B y = ( B A ) y
Hence, ( B A ) = g ( A B ) so ( B A ) A B . Defining u : Ω A × Ω B Ω A × Ω by u ( x , y ) = ( x , h ( y ) ) we have that:
u ( A B ) ( x , y ) = ( A B ) u 1 ( x , y ) = ( A B ) ( x , h 1 ( y ) ) = A x B h 1 ( y ) = A x h ( B ) y = A h ( B ) ( x , y )
It follows that A h ( B ) = u ( A B ) . Hence, A h ( B ) A B . □
Some results analogous to Theorem 3 hold for other entities.
Example 1.
We consider the simplest nontrivial example of a sequential product A B of observables. Let A = a 0 , a 1 , B = b 0 , b 1 be binary (diatomic) observables. Then, Ω A B = 0 , 1 × 0 , 1 and:
A B = a 0 b 0 , a 1 b 0 , a 0 b 1 , a 1 b 1
Except in trivial cases, A B has precisely the following nine parts to within equivalence:
A B , a 0 b 0 , a 1 + a 0 b 1 , a 1 b 0 , a 0 + a 1 b 1 , a 0 b 1 , a 1 + a 0 b 0 a 1 b 1 , a 0 + a 1 b 0 , a 0 b 0 + a 1 b 0 , a 0 b 1 + a 1 b 1 , a 0 , a 1 a 0 b 0 + a 1 b 1 , a 1 b 0 + a 0 b 1 , 1
Notice that the sixth of the parts is ( B A ) and the seventh is A as required by Theorem 3. Each of the parts is a function of A B . The parts listed correspond to the following functions f i : 0 , 1 × 0 , 1 1 , 2 , 3 , 4 , i = 1 , 2 , , 9 (Table 1).  □
Example 2.
Similarly to Example 1, for the two binary instruments I = I 0 , I 1 , J = J 0 , J 1 , we have the instrument I J with Ω I J = 0 , 1 × 0 , 1 and:
I J = I 0 J 0 , I 1 J 0 , I 0 J 1 , I 1 J 1
The nine parts of I J to within equivalence are:
I J , I 0 J 0 , I 0 J 1 + I 1 C J , I 1 J 0 , I 1 J 1 + I 0 C J I 0 J 1 , I 0 J 0 + I 1 C J , I 1 J 1 , I 1 J 0 + I 0 C J , C I J 0 , C I J 1 I 0 C J , I 1 C J , I 0 J 0 + I 1 J 1 , I 1 J 0 + I 0 J 1 , C I J
As in Example 1, the sixth part is ( J I ) , however, unlike the observable case, the seventh part is not I . In fact, unlike that case, I is not a part of ( I J ) .  □
If A O ( H ) , the corresponding Lüders instrument L A In ( H ) is defined by Ω L A = Ω A and L x A ( ρ ) = A x 1 / 2 ρ A x 1 / 2 for all ρ S ( H ) . It follows that [15]:
L X A ( ρ ) = x X A x 1 / 2 ρ A x 1 / 2
for all ρ S ( H ) , X Ω A . It is easy to check that ( L A ) = A . Hence, for B O ( H ) , we have that B L A if and only if B A .
Theorem 4.
(a) L A B = L A L B if and only if A x B y = B y A x for all x Ω A , y Ω B ; (b) ( L A B ) = ( L A L B ) = A B ; (c) An observable C satisfies C L A L B if and only if C A B .
Proof. 
(a) For all ρ S ( H ) , ( x , y ) Ω A × Ω B we have that:
( L A L B ) ( x , y ) ( ρ ) = L y B L x A ( ρ ) = L y B ( A x 1 / 2 ρ A x 1 / 2 ) = B y 1 / 2 A x 1 / 2 ρ A x 1 / 2 B y 1 / 2
On the other hand:
( L A B ) ( x , y ) ( ρ ) = ( A B ) ( x , y ) 1 / 2 ρ ( A B ) ( x , y ) 1 / 2 = ( A x B y ) 1 / 2 ρ ( A x B y ) 1 / 2 = ( A x 1 / 2 B y A x 1 / 2 ) 1 / 2 ρ ( A x 1 / 2 B y A x 1 / 2 ) 1 / 2
If A x B y = B y A x , we obtain:
( L A B ) ( x , y ) ( ρ ) = ( A x B y ) 1 / 2 ρ ( A x B y ) 1 / 2 = B y 1 / 2 A x 1 / 2 ρ A x 1 / 2 B y 1 / 2 = ( L A L B ) ( x , y ) ( ρ )
so that L A B = L A L B . Conversely, if L A B = L A L B , letting ρ = 1 n 1 where n = dim H , we obtain from (5) and (6) that:
B y A x = B y 1 / 2 A x B y 1 / 2 = A x 1 / 2 B y A x 1 / 2 = A x B y
It follows that B y A x = A x B y for all x Ω A , y Ω B [7]. (b) We already pointed out that ( L A B ) = A B . To show that ( L A L B ) = A B , applying (5) gives:
tr ρ ( L A L B ) ( x , y ) = tr ( L A L B ) ( x , y ) ( ρ ) = tr ( ρ A x 1 / 2 B y A x 1 / 2 ) tr ( ρ A x B y ) = tr ρ ( A B ) ( x , y )
Hence, ( L A L B ) = A B . (c) This follows from (b) and Theorem 1(a). □
Example 3.
We saw from Theorem 4(b) that ( L A L B ) = ( L A ) ( L B ) . We now show that ( I J ) I ^ J ^ in general. Let δ , γ S ( H ) and A , B O ( H ) . The instruments I x ( ρ ) = tr ( ρ A x ) δ and J y ( ρ ) = tr ( ρ B y ) γ are called trivial instruments with observables A , B and states δ , γ , respectively, [2]. We have that:
tr ( ρ I ^ x ) = tr I x ( ρ ) = tr tr ( ρ A x ) δ = tr ( ρ A x )
Hence, I ^ = A and similarly J ^ = B . For all ρ S ( H ) we obtain:
tr ρ ( I J ) ( x , y ) = tr ( I J ) ( x , y ) ( ρ ) = tr J y I x ( ρ ) = tr J y tr ( ρ A x ) δ = tr ( ρ A x ) tr J y ( δ ) = tr ( ρ A x ) tr tr ( δ B y ) γ = tr ( ρ A x ) tr ( δ B y )
On the other hand:
tr ρ I ^ x J ^ y = tr ( ρ A x B y )
Since the right hand sides of (7) and (8) are different in general, we conclude that ( I J ) I ^ J ^ .
We saw in Theorem 4(a) that L A B L A L B , in general. The following lemma shows they can differ in a striking way.
Lemma 2.
If A x = ϕ x ϕ x and B y = ψ y ψ y are atomic observables on H, then for all ρ S ( H ) , there exist numbers λ x y ( ρ ) 0 , 1 with x , y λ x y ( ρ ) = 1 such that L ( x , y ) A B ( ρ ) = λ x y ( ρ ) A x and ( L A L B ) ( x , y ) ( ρ ) = λ x y ( ρ ) B y for all ( x , y ) Ω A × Ω B .
Proof. 
For all ρ S ( H ) we have that:
( L A L B ) ( x , y ) ( ρ ) = L y B L x A ( ρ ) = B y A x ρ A x B y = ψ x ψ y ϕ x ϕ x ρ ϕ x ϕ x ψ y ψ y = ϕ x , ψ y 2 ϕ x , ρ ϕ x B y
Since:
A x B y A x = ϕ x ϕ x ψ y ψ y ϕ x ϕ x = ϕ x , ψ y 2 A x
we obtain:
( A x B y A x ) 1 / 2 = ϕ x , ψ y A x
Hence:
( L A B ) ( x , y ) ( ρ ) = ( A x B y A x ) 1 / 2 ρ ( A x B y A x ) 1 / 2 = ϕ x , ψ y 2 ϕ x , ρ ϕ x A x
Letting λ x y ( ρ ) = ϕ x , ψ y 2 ϕ x , ρ ϕ x , the result follows. □

4. Composite Systems

Let H 1 and H 2 be Hilbert spaces with dim H 1 = n 1 and dim H 2 = n 2 . If H 1 , H 2 represent quantum systems, we call H = H 1 H 2 a composite quantum system. For a E ( H ) , we define the reduced effects a 1 E ( H 1 ) , a 2 E ( H 2 ) by a 1 = 1 n 2 tr 2 ( a ) , a 2 = 1 n 1 tr 1 ( a ) . We view a i to be the effect a as measured in a system i = 1 , 2 . The map a a 1 is a surjective effect algebra morphism from E ( H ) onto E ( H 1 ) and similarly for a a 2 [6,7]. Conversely, if a E ( H 1 ) , b E ( H 2 ) , then a b E ( H ) and:
( a b ) 1 = 1 n 2 tr 2 ( a b ) = 1 n 2 tr ( b ) a
Similarly, ( a b ) 2 = 1 n 1 tr ( a ) b . It follows that:
( a 1 a 2 ) 1 = 1 n 2 tr ( a 2 ) a 1
and
( a 1 a 2 ) 2 = 1 n 1 tr ( a 1 ) a 2
An effect a E ( H ) is factorized if a = b c for b E ( H 1 ) , c E ( H 2 ) [2].
Lemma 3.
If a E ( H ) with a 0 , then a is factorized if and only if:
a = n 1 n 2 tr ( a ) a 1 a 2
Proof. 
If (9) holds, then a is factorized. Conversely, suppose a is factorized with a = b c , b E ( H 1 ) , c E ( H 2 ) . Then, a 1 = 1 n 2 tr ( c ) b and a 2 = 1 n 1 tr ( b ) c . Hence, b = n 2 tr ( c ) a 1 and c = n 1 tr ( b ) a 2 . We conclude that:
a = n 1 n 2 tr ( b ) tr ( c ) a 1 a 2 = n 1 n 2 tr ( a ) a 1 a 2
Corollary 1.
If a E ( H ) , then a = a 1 a 2 if and only if a = 0 or a = 1 .
Proof. 
If a = 0 or a = 1 , then clearly a = a 1 a 2 . Conversely, if a = a 1 a 2 , then by Lemma 3, a = 0 or tr ( a ) = n 1 n 2 . In the latter case, a = 1 . □
An effect is indecomposable if it has the form a = λ b where 0 λ 1 and b is an atom.
Theorem 5.
Let a E ( H ) be an atom a = P ψ where H = H 1 H 2 : (a) a is factorized if and only if a 1 and a 2 are indecomposable; (b) We can arrange the nonzero eigenvalues α 1 , α 2 , , α n of a 1 and the nonzero eigenvalues β 1 , β 2 , , β n of a 2 so that α i = n 1 n 2 β i , i = 1 , 2 , , n . Hence, if n 1 = n 2 , then the eigenvalues of a 1 and a 2 are identical.
Proof. 
The unit vector ψ H has a Schmidt decomposition ψ = i = 1 m λ i ψ i ϕ i , λ i 0 , λ i 2 = 1 . We have that:
a = ψ ψ = λ i ψ i ϕ i λ j ψ j ϕ j = i , j λ i λ j ψ i ϕ i ψ j ϕ j = i , j λ i λ j ψ i ψ j ϕ i ϕ j
Hence:
a 1 = 1 n 2 tr 2 ( a ) = 1 n 2 i , j λ i λ j tr 2 ψ i ψ j ϕ i ϕ j = 1 n 2 i , j λ i λ j δ i j ψ i ψ j = 1 n 2 λ i 2 P ψ i
and similarly:
a 2 = 1 n 1 λ i 2 P ϕ i
Now a is factorized if and only if ψ is factorized which is equivalent to m = 1 and ψ = ψ 1 ϕ 1 . Applying (10) and (11), we conclude that a is factorized if and only if a 1 = 1 n 2 λ 1 2 P ψ 1 and a 2 = 1 n 1 λ 1 2 P ϕ 1 , in which case a 1 and a 2 are indecomposable. This completes the proof of (a). To prove (b), we see from (10), (11) that the eigenvalues of a 1 , a 2 are α i = 1 n 2 λ i 2 and β i = 1 n 1 λ i 2 . It follows that α i = n 1 n 2 β i . □
If A O ( H 1 H 2 ) we define the reduced observables A 1 O ( H 1 ) , A 2 O ( H 2 ) by A 1 = A x 1 : x Ω A and A 2 = A x 2 : x Ω A . Note that A 1 ( A 2 ) is indeed an observable because:
x Ω A A x 1 = x Ω A 1 n 2 tr 2 ( A x ) = 1 n 2 tr 2 x Ω A A x = 1 n 2 tr 2 ( 1 1 1 2 ) = 1 1
Lemma 4.
If A O ( H 1 H 2 ) and ρ 1 S ( H 1 ) , then:
Φ ρ 1 A 1 = Φ ρ 1 1 2 / n 2 A
Proof. 
For X Ω A we have that:
Φ ρ 1 A 1 ( X ) = tr ( ρ 1 A x 1 ) = tr ρ 1 1 n 2 tr 2 ( A X ) = 1 n 2 tr ρ 1 tr 2 ( A X ) = 1 n 2 tr A X ( ρ 1 1 2 ) = tr ρ 1 1 n 2 1 2 A X = Φ ρ 1 1 2 / n 2 ( X )
The result now follows. □
In a similar way:
Φ ρ 2 A 2 = Φ 1 1 / n 1 ρ 2 A
For A O ( H 1 ) , we define the A-random measure on Ω A by
μ A ( X ) = 1 n 1 tr ( A X ) = tr 1 1 n 1 A X = Φ 1 1 / n 1 A ( X )
for all X Ω A . Thus, μ A is the distribution of A in the random state 1 1 / n 1 . If A 1 O ( H 1 ) , A 2 O ( H 2 ) , we define the composite observable:
B ( x , y ) = A 1 , x A 2 , y O ( H 1 H 2 )
In this case, Ω B = Ω A 1 × Ω A 2 and for Z Ω B , we have that:
B Z = ( x , y ) Z B ( x , y )
Hence, B X × Y = A 1 , X A 2 , Y .
Lemma 5.
B X × Y 1 = μ A 2 ( Y ) A 1 , X and B X × Y 2 = μ A 1 ( X ) A 2 , Y .
Proof. 
For x Ω A 1 , y Ω A 2 we obtain:
B ( x , y ) 1 = 1 n 2 tr 2 B ( x , y ) = 1 n 2 tr 2 ( A 1 , x A 2 , y ) = 1 n 2 tr ( A 2 , y ) A 1 , x
Hence:
B X × Y 1 = 1 n 2 tr ( A 2 , Y ) A 1 , X = μ A 2 ( Y ) A 1 , X
The second equation is similar. □
A transition probability from Ω 1 to Ω 2 is a map ν : Ω 1 × Ω 2 0 , 1 satisfying y Ω 2 ν x y = 1 for all x Ω 1 . (The matrix ν x y is called a stochastic matrix.) Let A O ( H 1 ) with outcome-space Ω 1 and let ν be a transition probability from Ω 1 to Ω 2 . Then, ( ν A ) y = x Ω 1 ν x y A x is an observable on H1 with outcome-space Ω 2 called a post-processing of A from Ω 1 to Ω 2 [12]. If we also have B O ( H 2 ) with outcome-space Ω 3 and μ a transition probability from Ω 3 to Ω 4 , we can form the post-processing μ • B.
Theorem 6.
(a) ( ν A ) ( μ B ) O ( H 1 H 2 ) with outcome-space Ω 2 × Ω 4 and is a post-processing α ( A B ) from Ω 1 × Ω 3 to Ω 2 × Ω 4 where α ( ( x , r ) , ( y , s ) ) = ν x y μ r s ; (b) If A O ( H 1 H 2 ) , then ( ν A ) 1 = ν A 1 and ( ν A ) 2 = ν A 2 .
Proof. 
(a) The map α : Ω 1 × Ω 3 Ω 2 × Ω 4 is a transition probability because α ( ( x , r ) , ( y , s ) ) 0 and:
( y , s ) Ω 2 × Ω 4 α ( ( x , r ) , ( y , s ) ) = y , s ν x y μ r s = 1
Moreover, ( ν A ) ( μ B ) O ( H 1 H 2 ) with outcome-space Ω 2 × Ω 4 and we have that:
[ ( ν A ) ( μ B ) ] ( y , s ) = ( ν A ) y ( μ B ) s = ( x Ω 1 ν x y A x ) ( r Ω 3 μ r s B r ) = x Ω 1 r Ω 3 ν x y μ r s A x B r = x , r α ( ( x , r ) , ( y , s ) ) A x B r = [ α ( A B ) ] ( y , s )
Hence, ( ν A ) ( μ B ) = α ( A B ) . (b) This follows from:
( ν A ) y 1 = x ν x y A x 1 = 1 n 2 tr 2 x ν x y A x = 1 n 2 x ν x y tr 2 ( A x ) = x ν x y A x 1 = ( ν A 1 ) y
That ( ν A ) 2 = ν A 2 is similar. □
We have seen in Theorem 2 that coexistence is equivalent to joint measurability. This is used in the next theorem [13].
Theorem 7.
(a) If A 1 , B 1 O ( H 1 ) coexist with joint observable C 1 and A 2 , B 2 coexist with joint observable C 2 , then A 1 A 2 , B 1 B 2 coexist with joint observable C = C 1 C 2 ; (b) If A , B O ( H 1 H 2 ) coexist with joint observable C, then A 1 , B 1 coexist with joint observable C 1 and A 2 , B 2 coexist with joint observable C 2 .
Proof. 
(a) We write C 1 , ( x , y ) for ( x , y ) Ω A 1 × Ω B 1 and C 2 , ( x , y ) for ( x , y ) Ω A 2 × Ω B 2 . Then:
C ( x , y , x , y ) = C 1 , ( x , y ) C 2 , ( x , y )
and we have that:
( y , y ) C ( x , y , x , y ) = y C 1 , ( x , y ) y C 2 , ( x , y ) = A 1 , x A 2 , x = ( A 1 A 2 ) ( x , x )
Moreover:
( x , x ) C ( x , y , x , y ) = x C 1 , ( x , y ) x C 2 , ( x , y ) = B 1 , y B 2 , y = ( B 1 B 2 ) ( y , y )
and the result follows. (b) For all ( x , y ) Ω A × Ω B , we obtain:
A x 1 = y C ( x , y ) 1 = 1 n 2 tr 2 y C ( x , y ) = y 1 n 2 tr ( C ( x , y ) ) = y C ( x , y ) 1
Similarly, B y 1 = x C ( x , y ) 1 so A 1 , B 1 coexist with joint observable C 1 . The result for A 2 , B 2 is similar. □
For an instrument I In ( H 1 H 2 ) on the composite system, the reduced instrument on system 1 is defined by [9,10]
I x 1 ( ρ 1 ) = 1 n 2 tr 2 I x ( ρ 1 1 2 )
for all ρ 1 S ( H 1 ) , x Ω I . Similarly:
I x 2 ( ρ 1 ) = 1 n 1 tr 1 I x ( 1 1 ρ 2 )
for all ρ 2 S ( H 2 ) , x Ω I .
Theorem 8.
( I 1 ) = ( I ^ ) 1 and ( I 2 ) = ( I ^ ) 2 .
Proof. 
For all ρ 1 S ( H 1 ) , we have that:
tr ρ 1 ( I ^ ) x 1 = 1 n 2 tr ρ 1 tr 2 ( I ^ ) x = 1 n 2 tr ( ρ 1 1 2 ) I ^ x = 1 n 2 tr I x ( ρ 1 1 2 ) = tr I x 1 ( ρ 1 ) = tr ρ 1 ( I 1 ) x
We conclude that ( I 1 ) = ( I ^ ) 1 and similarly, ( I 2 ) = ( I ^ ) 2 . □
For I In ( H 1 ) we define the I -random measure on Ω I by
μ I ( X ) = 1 n 1 tr I X ( 1 1 )
For I 1 In ( H 1 ) , I 2 In ( H 2 ) we define J = I 1 I 2 In ( H 1 H 2 ) with outcome-space Ω I 1 × Ω I 2 by J ( x , y ) = I 1 , x I 2 , y . It is easy to check that J is indeed an instrument.
Theorem 9.
Let J = I 1 I 2 In ( H 1 H 2 ) . (a) J ( x , y ) 1 ( ρ 1 ) = μ I 2 ( y ) I 1 , x ( ρ 1 ) for all ρ 1 S ( H 1 ) and J ( x , y ) 2 ( ρ 2 ) = μ I 1 ( x ) I 2 , y ( ρ 2 ) for all ρ 2 S ( H 2 ) . (b) ( I 1 I 2 ) = I ^ 1 I ^ 2 .
Proof. 
(a) For all ρ 1 S ( H 1 ) we have that:
J ( x , y ) 1 ( ρ 1 ) = 1 n 2 tr 2 J ( x , y ) ( ρ 1 1 2 ) = 1 n 2 tr 2 I 1 , x I 2 , y ( ρ 1 2 ) = 1 n 2 tr 2 I 1 , x ( ρ 1 ) I 2 , y ( 1 2 ) = 1 n 2 tr I 2 , y ( 1 2 ) I 1 , x ( ρ 1 ) = μ I 2 ( y ) I 1 , x ( ρ 1 )
Similarly, J ( x , y ) 2 ( ρ 2 ) = μ I 1 ( x ) I 2 , y ( ρ 2 ) for all ρ 2 S ( H 2 ) . (b) For all ρ 1 S ( H 1 ) , ρ 2 S ( H 2 ) we have that:
tr ρ 1 ρ 2 ( I 1 I 2 ) ( x , y ) = tr I 1 , x I 2 , y ( ρ 1 ρ 2 ) = tr I 1 , x ( ρ 1 ) I 2 , y ( ρ 2 ) = tr I 1 , x ( ρ 1 ) tr I 2 , y ( ρ 2 ) = tr ρ 1 I ^ 1 , x tr ρ 2 I ^ 2 , y = tr ρ 1 ρ 2 ( I ^ 1 , x I ^ 2 , y )
and the result follows. □
A Kraus instrument is an instrument of the form I x ( ρ ) = S x ρ S x * where x S x * S x = 1 , x Ω I . The operators S x are called Kraus operators for I [5].
Lemma 6.
Let I 1 In ( H 1 ) , I 2 In ( H 2 ) be Kraus instruments with operators S 1 , x , S 2 , y , respectively. (a) J = I 1 I 2 is a Kraus instrument with operators S 1 , x S 2 , y ; (b) J 1 , J 2 are Kraus instruments with operators:
T ( x , y ) = 1 n 2 tr ( S 2 , y S 2 , y * ) 1 / 2 S 1 , x R ( x , y ) = 1 n 1 tr ( S 1 , x S 1 , x * ) 1 / 2 S 2 , y
Proof. 
(a) For all ρ 1 S ( H 1 ) , ρ 2 S ( H 2 ) we have that:
J ( x , y ) ( ρ 1 ρ 2 ) = ( I 1 , x × I 2 , y ) ( ρ 1 ρ 2 ) = I 1 , x ( ρ 1 ) I 2 , y ( ρ 2 ) = S 1 , x ρ 1 S 1 , x * S 2 , y ρ 2 S 2 , y * = S 1 , x S 2 , y ( ρ 1 ρ 2 ) S 1 , x * S 2 , y *
and the result follows. (b) For ρ S ( H 1 ) we obtain:
J ( x , y ) 1 ( ρ 1 ) = 1 n 2 tr I 2 , y ( 1 2 ) I 1 , x ( ρ 1 ) = 1 n 2 tr ( S 2 , y S 2 , y * ) S 1 , x ρ 1 S 1 , x *
This can be considered to be a Kraus instrument with operators T ( x , y ) given above. The result for J 2 is similar. □
Notice that a Lüders instrument defined by L x A ( ρ 1 ) = A x 1 / 2 ρ 1 A x 1 / 2 for all ρ 1 S ( H 1 ) is a particular case of a Kraus instrument with operators A x 1 / 2 [15].
Corollary 2.
Let A O ( H 1 ) , B O ( H 2 ) : (a) L x A L y B = L ( x , y ) A B ; (b) ( L x A L y B ) 1 = L ( x , y ) C where C = 1 n 2 tr ( B y ) A x and ( L x A L y B ) 2 = L ( x , y ) D where D = 1 n 2 tr ( A x ) B y .
We say that a Kraus instrument I In ( H 1 H 2 ) with operators R x is factorized if R x = S x T x for all x Ω I . We conjecture that if I In ( H 1 H 2 ) is Kraus, then I 1 and I 2 need not be Kraus. However, we do have the following result.
Lemma 7.
If I In ( H 1 H 2 ) is Kraus and factorized, then I 1 and I 2 are Kraus.
Proof. 
If the operators R x for I satisfy R x = S x T x , then for all ρ 1 S ( H 1 ) , we have that:
I x 1 ( ρ 1 ) = 1 n 2 tr 2 I x ( ρ 1 1 2 ) = 1 n 2 tr 2 R x ( ρ 1 1 2 ) R x * = 1 n 2 tr 2 S x T x ( ρ 1 1 2 ) S x * T x * = 1 n 2 tr 2 ( S x ρ 1 S x * T x T x * ) = 1 n 2 tr ( T x T x * ) S x ρ 1 S x *
Hence, I 1 is Kraus with operators 1 n 2 tr ( T x T x * ) 1 / 2 S x . Similarly, I 2 is Kraus with operators 1 n 1 tr ( S x S x * ) 1 / 2 T x . □
We do not know if the converse of Lemma 7 holds. We now consider trivial instruments (see Example 3).
Lemma 8.
Let I 1 In ( H 1 ) , I 2 In ( H 2 ) be trivial instruments with:
I 1 , x ( ρ 1 ) = tr ( ρ 1 A x ) α , I 2 , y ( ρ 2 ) = tr ( ρ 2 B y ) β
(a) I 1 I 2 In ( H 1 H 2 ) is trivial with observable A B and state α β (b) ( I 1 I 2 ) 1 , ( I 1 I 2 ) 2 are trivial with observables μ B ( y ) A x , μ A ( x ) B y and states α , β , respectively.
Proof. 
(a) For all ( x , y ) Ω I 1 × Ω I 2 , ρ 1 S ( H 1 ) , ρ 2 S ( H 2 ) , we have that:
( I 1 I 2 ) ( x , y ) ( ρ 1 ρ 2 ) = I 1 , x ( ρ 1 ) I 2 , y ( ρ 2 ) = tr ( ρ 1 A x ) α tr ( ρ 2 B y ) β = tr ( ρ 1 A x ) tr ( ρ 2 B y ) α β = tr ( ρ 1 A x ρ 2 B y ) α β = tr ( ρ 1 ρ 2 A B ( x , y ) ) α β
The result now follows. (b) This follows from:
( I 1 I 2 ) ( x , y ) 1 ( ρ 1 ) = 1 n 2 tr 2 I 1 , x I 2 , y ( ρ 1 1 2 ) = 1 n 2 tr 2 I 1 , x ( ρ 1 ) I 2 , y ( 1 2 ) = 1 n 2 tr I 2 , y ( 1 2 ) I 1 , x ( ρ 1 ) = 1 n 2 tr tr ( 1 2 B y ) β tr ( ρ 1 A x ) α = 1 n 2 tr ( B y ) tr ( ρ 1 A x ) α = tr ρ 1 μ B ( y ) A x α
and similarly:
( I 1 I 2 ) ( x , y ) 2 ( ρ 2 ) = tr ρ 2 μ A ( x ) B y β
Lemma 9.
Let I In ( H 1 H 2 ) be trivial with I x ( ρ ) = tr ( ρ A x ) α . (a) I 1 , I 2 are trivial with observables A x 1 , A x 2 and states tr 2 ( α ) , tr 1 ( α ) , respectively; (b) Letting J = I 1 I 2 we have that J is trivial with observable A 1 A 2 and state tr 2 ( α ) tr 1 ( α ) . Moreover, J ( x , y ) 1 = I x 1 and J ( x , y ) 2 = I y 2 for all ( x , y ) Ω J .
Proof. 
(a) For all ρ 1 S ( H 1 ) and x Ω I , we have that:
I x 1 ( ρ 1 ) = 1 n 2 tr 2 I x ( ρ 1 1 2 ) = 1 n 2 tr 2 tr ( ρ 1 1 2 ) A x α = 1 n 2 tr ( ρ 1 1 2 ) A x tr 1 ( α ) = 1 n 2 tr tr 2 ( A x ) ρ 1 tr 2 ( α ) = tr ρ 1 1 n 2 tr 2 ( A x ) tr 2 ( α ) = tr ( ρ 1 A x 1 ) tr 2 ( α )
Similarly, I x 2 ( ρ 2 ) = tr ( ρ 2 A x 2 ) tr 1 ( α ) so the result follows. (b) This result follows from Lemma 8(b). □
We now consider M M s for composite systems. A single probe M M on H = H 1 H 2 has the form M = ( H , K , η , ν , F ) as defined before. As discussed earlier, M ^ In ( H ) is the instrument measured by M . Then, M ^ 1 In ( H 1 ) and for ρ 1 S ( H 1 ) , we obtain:
M ^ x 1 ( ρ 1 ) = 1 n 2 tr 2 M ^ x ( ρ 1 1 2 ) = 1 n 2 tr 2 tr K ν ( ρ 1 1 2 η ) ( 1 1 1 2 F x )
We have a similar expression for M ^ 2 In ( H 2 ) .
Corresponding to M , we define the reduced M M M 1 = ( H 1 , K , η , ν 1 , F ) where ν 1 S ( H 1 K ) is given by
ν 1 ( ρ 1 η ) = 1 n 2 tr 2 ν ( ρ 1 1 2 η )
We then have for ρ 1 S ( H 1 ) that:
M ^ 1 , x ( ρ 1 ) = tr K ν 1 ( ρ 1 η ) ( 1 1 F x ) = 1 n 2 tr K tr 2 ν ( ρ 1 1 2 η ) ( 1 1 F x )
Similarly, we define M 2 = ( H 2 , K , η , ν 2 , F ) and an analogous formula for M ^ 2 . Notice that (12) and (13) are quite similar and they are essentially an interchange of the two partial traces. We now show that they coincide.
Theorem 10.
(a) Let H 1 , H 2 , H 3 be finite-dimensional Hilbert spaces and let A L ( H 1 H 2 H 3 ) , B L ( H 3 ) . Then:
tr 2 tr 3 A ( 1 1 1 2 B ) = tr 3 tr 2 ( A ) ( 1 1 B )
(b) M ^ 1 = M ^ 1 and M ^ 2 = M ^ 2 .
Proof. 
(a) First suppose that A = A 1 A 2 A 3 is factorized. We then obtain:
tr 2 tr 3 A ( 1 1 1 2 B ) = tr 2 tr 3 A 1 A 2 A 3 ( 1 1 1 2 B ) = tr 2 tr 3 ( A 1 A 2 A 3 B ) = tr 2 A 1 A 2 tr ( A 3 B ) = tr ( A 3 B ) tr 2 ( A 1 A 2 ) = tr ( A 3 B ) tr ( A 2 ) A 1 = tr ( A 2 ) tr 3 ( A 1 A 3 B ) = tr 3 tr ( A 2 ) ( A 1 A 3 ) ( 1 1 B ) = tr 3 tr 2 ( A 1 A 2 A 3 ) ( 1 1 B ) = tr 3 tr 2 ( A ) ( 1 1 B )
Hence, (14) holds when A is factorized. Since any A L ( H 1 H 2 H 3 ) is a linear combination of factorized operators, (14) holds in general. (b) Letting A = ν ( ρ 1 1 2 η ) , B = F x and K = H 3 in (14), we conclude that (12) and (13) coincide. Hence, M ^ 1 = M ^ 1 and similarly, M ^ 2 = M ^ 2 . □
We considered single probe composite M M s. We now briefly discuss general composite M M s. Let M i = ( H i , K i , η i , ν i , F i ) , i = 1 , 2 be two M M s. Define the unitary swap operator [2]:
U : H 1 H 2 K 1 K 2 H 1 K 1 H 2 K 2
by
U ( ϕ 1 ϕ 2 ψ 1 ψ 2 ) = ϕ 1 ψ 1 ϕ 2 ψ 2
We now define the channel ν 1 ν 2 C ( H 1 H 2 K 1 K 2 ) by
ν 1 ν 2 ( ρ 1 ρ 2 η 1 η 2 ) = U * ν 1 ( ρ 1 η 2 ) ν 2 ( ρ 2 η 2 ) U
The composite of M 1 and M 2 is declared to be:
M = M 1 M 2 = ( H 1 H 2 , K 1 K 2 , η 1 η 2 , ν 1 ν 2 , F 1 F 2 )
For ρ S ( H 1 H 2 ) , we have that:
M ^ ( x , y ) ( ρ ) = tr K 1 K 2 ν 1 ν 2 ( ρ η 1 η 2 ) ( 1 1 1 2 F 1 , x F 2 , y )
The next result shows that M has desirable properties.
Theorem 11.
(a) For ρ = ρ 1 ρ 2 S ( H 1 H 2 ) we have:
M ^ ( x , y ) ( ρ ) = M ^ 1 , x ( ρ 1 ) M ^ 2 , y ( ρ 2 )
(b) defining M ^ 1 and M ^ 2 in the usual way, we obtain:
M ^ ( x , y ) 1 ( ρ 1 ) = 1 n 2 tr M ^ 2 , y ( 1 2 ) M ^ 1 , x ( ρ 1 )
and:
M ^ ( x , y ) 2 ( ρ 2 ) = 1 n 1 tr M ^ 1 , x ( 1 1 ) M ^ 2 , y ( ρ 2 )
for all ρ 1 S ( H 1 ) , ρ 2 S ( H 2 ) .
Proof. 
(a) Applying (15) and (16) we obtain:
M ^ ( x , y ) ( ρ ) = tr K 1 K 2 U * ν 1 ( ρ 1 η 1 ) ν 2 ( ρ 2 × η 2 ) U ( 1 1 1 2 F 1 , x F 2 , y ) = tr K 1 K 2 ν 1 ( ρ 1 η 1 ) ν 2 ( ρ 2 η 2 ) U ( 1 1 1 2 F 1 , x F 2 , y ) U * = tr K 1 tr K 2 ν 1 ( ρ 1 η 1 ) ν 2 ( ρ 2 η 2 ) ( 1 1 F 1 , x 1 2 F 2 , y ) = tr K 1 tr K 2 ν 1 ( ρ 1 η 1 ) ( 1 1 F 1 , x ) ν 2 ( ρ 2 η 2 ) ( 1 2 F 2 , y ) = M ^ 1 , x ( ρ 1 ) M ^ 2 , y ( ρ 2 )
(b) For ρ 1 S ( H 1 ) we have that:
M ( x , y ) 1 ( ρ 1 ) = 1 n 2 tr 2 M ^ ( x , y ) ( ρ 1 1 2 ) = 1 n 2 tr 2 M ^ 1 , x ( ρ 1 ) M ^ 2 , y ( 1 2 ) = 1 n 2 tr M ^ 2 , y ( 1 2 ) M ^ 1 , x ( ρ 1 )
The expression for M ^ 2 is similar. □

5. Concluding Remarks

In this article, we only considered finite-dimensional Hilbert spaces. One reason for this was to avoid various measure theoretic details and thus greatly simplify the exposition. A second reason was that the direction of quantum investigations has largely changed over the last twenty years. This modern research is mainly concerned with more practical matters involving quantum computation and information theory [2,3,4,13]. Although it is restrictive to only consider finite-dimensional Hilbert spaces, the resulting structures are general enough to include these modern theories. Nevertheless, with more work, many of our results extend to the infinite-dimensional case. For example, in this situation, an observable is defined as an effect-valued measure A : B ( R ) E ( H ) , where B ( R ) is the collection of Borel subsets of R [1,2,3]. If B : B ( R ) E ( H ) is another observable, we can define their sequential product A B : B ( R 2 ) E ( H ) by
A B ( X × Y ) = A ( X ) B ( Y )
for all X , Y B ( R ) . Some measure theoretic details are required to show that A B as defined in (17) for product sets X × Y extends to an effect-valued measure on B ( R 2 ) . In a similar way, we define an instrument as an operation-valued measure on B ( R ) . Just as we did before, an instrument I measures a unique observable I ^ . Moreover, it is straightforward to define measurement models for infinite-dimensional Hilbert spaces. One can define parts of observables in a natural way and employing measure theory, Theorem 1 extends to infinite-dimensions. We leave it to the reader to check which other theorems extend.
We now summarize the main results of this article. The basic concept of our work is an effect a E ( H ) . An effect represents the simplest type of experiment in which the result of a measurement is either yes or no (also called true or false). An example would be to flip a switch: a light either goes on or does not go on. An observable A corresponds to a more complicated experiment that has a finite number of possible outcomes x Ω A . If a measurement of A results in outcome x, then the effect A x is true (has answer yes) and otherwise A x is false (has answer no). If our underlying physical system is in a state ρ S ( H ) , then the probability that a measurement of A results in outcome x is given by Born’s rule tr ( ρ A x ) . Various physical apparatuses may be employed to measure an observable A. The most common of these is I = I x : x Ω A where I x : S ( H ) S ( H ) is a completely positive linear map. We say that I measures the observable A if tr I x ( ρ ) = tr ( ρ A x ) . Thus, I and A have the same probability distribution. However, I gives more information than its unique measured observable I ^ = A because I determines the updated state I x ( ρ ) / tr ( ρ A x ) when A has outcome x.
Finally, we have the concept of a finite measurement model ( M M ) given by M = ( H , K , η , ν , F ) . In this case, H is the Hilbert space describing the system being observed and K is the Hilbert space describing the measuring apparatus which is in the initial state η S ( K ) . In order to perform the measurement, the two systems interact, which is described by the tensor product H K and a channel ν on H K . The initial state ρ S ( H ) of the observed system combines with η to form the state ρ η H K . This state is sent through the channel ν and the probe observable F O ( K ) is applied to give a measurement outcome. The model M measures a unique instrument M ^ In ( H ) given by
M ^ x ( ρ ) = tr K ν ( ρ η ) ( I F x )
and a unique observable M x O ( H ) .
We defined a function A = f ( B ) of an observable B and call A a part of B. We can then think of A as being obtained from B by piecing together parts of B. This concept is extended to parts of instruments and M M ’s. Theorem 1 shows that parts are preserved under measurements. For example, f ( I ^ ) = f ( I ) for all instruments I and f ( M ^ ) = f ( M ) for all M M ’s M . We then consider the coexistence of observables and Theorem 2 shows that observables coexist if and only if they are jointly measurable. We next introduce the concepts of sequential products A B and conditioning ( B A ) of observables A , B O ( H ) . We interpret A B as the observable obtained by first measuring A and then measuring B and ( B A ) as the observable B conditioned on first measuring the observable A. Theorem 3 shows that A , ( B A ) and A f ( B ) are parts of A B . Sequential products of instruments are also discussed. For an observable A O ( H ) , we consider the corresponding Lüders instrument L A In ( H ) . Theorem 4 shows that L A B = L A L B if and only if A and B commute. Moreover, ( L A B ) = A B and C is a part of L A B if and only if C is a part of A B .
Section 4 considers composite systems H = H 1 H 2 . If a E ( H ) , we define the reduced effects a 1 E ( H 1 ) , a 2 E ( H 2 ) , where a i is thought of as the effect as measured in the system H i , i = 1 , 2 . This concept is extended to the reduced observables A i O ( H i ) , i = 1 , 2 , for A O ( H 1 H 2 ) . Theorem 6 proves results concerning post-processing for composite systems and Theorem 7 shows that coexistence is preserved on taking tensor products and under reducing observables. Reduced instruments I i In ( H i ) , i = 1 , 2 , for I In ( H 1 H 2 ) are also defined and Theorems 8 and 9 show (among other things) that ( I 1 ) = ( I ^ ) 1 , ( I 2 ) = ( I ^ ) 2 and ( I 1 I 2 ) = I ^ 1 I ^ 2 . Finally, we consider M M ’s for composite systems. A single probe M M on H = H 1 H 2 has the usual form M = ( H , K , η , ν , F ) . Then, for ρ 1 S ( H 1 ) , we obtain M ^ 1 In ( H 1 ) given by
M ^ x 1 ( ρ 1 ) = 1 n 2 tr 2 tr K η ( ρ 1 1 2 η ) ( 1 1 1 2 F x )
where n 2 = dim H 2 . There is a similar expression for M ^ 2 In ( H 2 ) . Corresponding to M , we define the reduced M M given by M 1 = ( H 1 , K , η , ν 1 , F ) where ν 1 S ( H 1 K ) is given by
ν 1 ( ρ 1 η ) = 1 n 2 tr 2 ν ( ρ 1 1 2 η )
We think of M 1 as the M M M as measured by the H 1 system. We then obtain:
M ^ 1 , x ( ρ 1 ) = 1 n 2 tr K tr 2 ν ( ρ 1 1 2 η ) ( 1 1 F x )
Similarly, we define M 2 = ( H 2 , K , η , ν 2 , F ) and obtain an analogous formula for M ^ 2 . Now, M ^ x 1 and M ^ 1 , x essentially represent the same object and (18), (19) are quite similar. In fact, Theorem 10 shows that they coincide. This paper closes with a definition of a general composite MM and Theorem 11 shows that this definition has desirable properties.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Busch, P.; Grabowski, M.; Lahti, P. Operational Quantum Physics; Springer: Heidelberg/Berlin, Germany, 1995. [Google Scholar]
  2. Heinosaari, T.; Ziman, M. The Mathematical Language of Quantum Theory; Cambridge University Press: Cambridge, UK, 2012. [Google Scholar]
  3. Nielson, M.; Chuang, I. Quantum Computation and Quantum Information; Cambridge University Press: Cambridge, UK, 2000. [Google Scholar]
  4. Fillipov, S.; Heinosaari, T.; Leppäjärvi, L. Simulability of observables in general probabilistic theories. Phys. Rev. 2018, A97, 062102. [Google Scholar] [CrossRef] [Green Version]
  5. Kraus, K. States, Effects and Operations; Springer: Heidelberg/Berlin, Germany, 1983. [Google Scholar]
  6. Gudder, S.; Greechie, R. Sequential Products on effect algebras. Rep. Math. Phys. 2002, 49, 87–111. [Google Scholar] [CrossRef]
  7. Gudder, S.; Nagy, G. Sequential quantum measurements. J. Math. Phys. 2001, 42, 5212–5222. [Google Scholar] [CrossRef]
  8. Gudder, S. Conditioned observables in quantum mechanics. arXiv 2020, arXiv:quant-ph 2005.04775. [Google Scholar]
  9. Gudder, S. Quantum instruments and conditioned observables. arXiv 2020, arXiv:quant-ph 2005.08117. [Google Scholar]
  10. Gudder, S. Finite quantum instruments. arXiv 2020, arXiv:quant-ph 2005.13642. [Google Scholar]
  11. Ozawa, M. Operations, disturbance, and simultaneous measurability. Phys. Rev. 2001, A63, 032109. [Google Scholar] [CrossRef] [Green Version]
  12. Heinosaari, T.; Reitzner, D.; Stano, R.; Ziman, M. Coexistence of quantum operations. J. Phys. 2009, A42, 365302. [Google Scholar] [CrossRef]
  13. Heinosaari, T.; Miyandera, T.; Reitzner, D. Strongly incompatible quantum devices. Found. Phys. 2014, 44, 34–57. [Google Scholar] [CrossRef] [Green Version]
  14. Lahti, P. Coexistence and joint measurability in quantum mechanics. Int. J. Theor. Phys. 2003, 42, 893–906. [Google Scholar] [CrossRef]
  15. Lüders, G. Über due Zustandsänderung durch den Messprozess. Ann. Physik 1951, 6, 322–328. [Google Scholar]
Table 1. Function values.
Table 1. Function values.
Function ( 0 , 0 ) ( 0 , 1 ) ( 1 , 0 ) ( 1 , 1 )
f 1 1234
f 2 1222
f 3 2212
f 4 2122
f 5 2221
f 6 1212
f 7 1122
f 8 1221
f 9 1111
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gudder, S.P. Parts and Composites of Quantum Systems. Symmetry 2021, 13, 1031. https://doi.org/10.3390/sym13061031

AMA Style

Gudder SP. Parts and Composites of Quantum Systems. Symmetry. 2021; 13(6):1031. https://doi.org/10.3390/sym13061031

Chicago/Turabian Style

Gudder, Stanley P. 2021. "Parts and Composites of Quantum Systems" Symmetry 13, no. 6: 1031. https://doi.org/10.3390/sym13061031

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop