Operator Algebras and Their Application in Physics: Lecture Notes On
Operator Algebras and Their Application in Physics: Lecture Notes On
APPLICATION IN PHYSICS
TU Bergakademie Freiberg
Institute for Geophysics
Geomathematics and Geoinformatics Group
Preamble
Around the year 1927 physicist, lead by Heisenberg and Schrödinger, discovered a new
theory for subatomic particles, called quantum mechanics. This theory radically differed
from classical mechanics in its physical assumptions as well as in its mathematical formu-
lation. The formal developments of quantum mechanics rapidly made it one of the most
successful physical theories. Nevertheless, its physical foundations were very obscure and
unjustifiable, and its mathematical basis were often shaky. To provide a clear physical
justification and a rigorous mathematical foundation for the theory, several mathemati-
cians, lead by von Neumann, Weyl, and Jordan, set out to define an algebraic approach
for quantum mechanics. In this lecture we follow their path and see how the physi-
cal formulation of the standard Schrödinger quantum mechanics appears from general
physical arguments, paired with the brilliant observations of Heisenberg, and impeccable
mathematics of von Neumann and friends...
In the first part of the lecture, we will develop the basic theory of operator algebras.
Here we will not make any connections to physics and treat the subject in its pure form. In
particular, we discuss the function calculus and the Gelfand-Naimark-Segal construction
of representations for C ‹ -algebras. These tools will give us the necessary mathematical
background to deal with C ‹ -algebras when they appear in physics.
In the second part of the lecture we establish a connection between pure mathe-
matics and physics. We will see how general physical principles lead us to an algebraic
description of physical theories. Combining this description with the Heisenberg’s uncer-
tainty principle and the canonical commutation relations developed by Born and Jordan
we end up with the C ‹ -algebra that describes a quantum mechanical particle. The fi-
nal piece of the puzzle follows from the Stone-von Neumann theorem that justifies the
almost exclusive use of Schrödinger’s formulation for quantum mechanics.
2
Contents
1 Operator algebras: Banach and C ‹ -algebras 4
1.1 Basic properties of operator algebras . . . . . . . . . . . . . . . . . . . . 4
1.2 Function calculus for C ‹ -algebras . . . . . . . . . . . . . . . . . . . . . . 18
1.3 The GNS construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.3.1 Order structure on the C ‹ -algebra . . . . . . . . . . . . . . . . . . 24
1.3.2 States of the C ‹ -algebra . . . . . . . . . . . . . . . . . . . . . . . 27
1.3.3 Representations of a C ‹ -algebra . . . . . . . . . . . . . . . . . . . 31
3
1 Operator algebras: Banach and C ‹-algebras
Section remarks: This section is based on the volume 1 and 3 of the series by Kadison
and Ringrose [1, 2]. Some of the reminders and definitions that concern more standard
statements of functional analysis are mostly from [3]. Further reading: for the abstract
approach to C ‹ -algebras and von Neumann algebras the book by Sakai [4] is great but
is a little more advanced; another very satisfying read is the book by Pedersen [5]; the
four volumes of the aforementioned Kadison and Ringrose bible are anyway a treat for
the soul: the books cover most of the topics of old fashioned C ‹ -algebras and develop
the theory very slowly and self-consistently.
In a nutshell, an algebra is a vector space in which we can multiply vectors. More formally
we define the algebra as follows.
Definition 1.1 (Algebra). An algebra A over a field K “ pR, Cq is a vector space over K,
equipped with multiplication that
i ) distributes over addition (from left and right):
f ¨ pg ` hq “ f ¨ g ` f ¨ h, pf ` gq ¨ h “ f ¨ h ` g ¨ h; (1)
ii ) is associative:
pf ¨ gq ¨ h “ f ¨ pg ¨ hq; (2)
Definition 1.2. An algebra A over a field K is a vector space, whose underlying additive
group has a ring structure compatible with multiplication with scalars.
To recall the definitions of the vector spaces and rings consult the following remarks.
4
Reminder: Vector space. A vector space over a field K is a set of points V equipped
with vector addition ` : V ˆ V Ñ V and scalar multiplication ¨ : F ˆ V Ñ V . These
operations have to satisfy the following requirements.
For f, g P V and α, β P K the scalar multiplication has to
i ) distribute over the vector addition: αpf ` gq “ αf ` αg,
ii ) distribute over the field addition: pα ` βqf “ αf ` βf ,
iii ) be compatible with the field multiplication: pαβqf “ αpβf q,
iv ) has the neutral element: 1f “ f .
The vector addition has to be commutative and V with ` : V Ñ V must be a
commutative group. That is for f, g, h P V and α, β P K the addition
i ) is commutative (aka. Abelian): f ` g “ g ` f ,
ii ) is associative: pf ` gq ` h “ f ` pg ` hq,
iii ) has a neutral element e P V : e ` f “ f
iv ) is such that for every f P V there is an inverse f ´1 P V : f ` f ´1 “ e.
Example 1.3. To visualize the concept let us recall some algebras that we know and love.
i ) The simplest example of an algebra is the algebraic field K itself.
ii ) A more elaborate example is the space Mpn, Cq of n ˆ n matrices with complex
coefficients. This algebra is unital. The unit being the identity matrix.
iii ) Yet another example is the space R ˆ R with the Hadamard product defined by
p1, 0q ¨ p1, 0q “ p1, 0q, p0, 1q ¨ p0, 1q “ p0, 1q, p1, 0q ¨ p0, 1q “ p0, 0q, (4)
5
iv ) Finally, consider the space of all polynomials in a single variable. We will call this
space P l. It is a vector space, and equipped with the pointwise product it becomes
a unital algebra, whose unit is the constant 1.
As we will see, Banach algebras can be quite small (that is contain few elements).
Nevertheless, they are very general, as they have only little structure compared to the
algebras we will primarily use. We define this additional structure as follows.
Reminder: Adjoint. Let BpHq be the space of bounded operators on the Hilbert
space H, and let x¨, ¨y denote the inner product on H. Then, for A P BpHq we define
its adjoint as the operator A: P BpHq such that
It is not difficult to see that the adjoint on the space of bounded linear operators on
a Hilbert space satisfies all the requirements of Definition 1.5. Indeed, the adjoint is
the reason why we made these requirements in the first place. However, in the case of
bounded linear operators, we define an adjoint using the scalar product and then check
that it satisfies the properties in the Definition. For algebras, the logic is reversed. We
don’t have the scalar product at our disposal, and so we require the properties as the
defining features of the mapping. In this sense the involution, sometimes simply called
the star (operation), generalizes the concept of the adjoint for bounded linear operators.
With this additional structure we can now define the algebras, that will accompany
us for the rest of the lecture.
6
Definition 1.7. A C ‹ -algebra A is a complex Banach ‹-algebra that has the C ‹ -property:
In the definition of Banach ‹-algebras some authors do not require the involution to
be isometric. In our case, this difference will not play a big role, because we will only
use either Banach algebras (without the star), or C ‹ -algebras. For the latter, however,
the involution is always isometric: since if A is an element in the C ‹ -algebra A then the
C ‹ -property implies
and thus }A} ď }A‹ }. Upon replacing A Ñ A‹ we get the inverse inequality and thus
}A} “ }A‹ }. Notice, however, that the converse is not true: an isometric involution is
not enough to ensure the C ‹ -property. Thus, in general C ‹ -algebras have more structure
than Banach ‹-algebras.
The name giving for the elements of the Banach ‹-algebra follows the usual conven-
tion. We introduce the standard notation.
pI ‹ Iq‹ “ pI ‹ q‹ “ I “ I ‹ I “ I ‹ . (8)
7
Proof. Calculate
and
Before we look at some examples of C ‹ -algebras, let us quickly refresh some basic
definitions in topology.
Proof. Clearly, Cb pXq is a vector space. Since a pointwise product of two continuous
functions is again a continuous function, the space Cb pXq is algebraically closed. Since
the uniform limit of continuous functions is continuous, the space Cb pXq is complete.
Hence, Cb pXq with the supremum norm and the pointwise product is a Banach algebra.
It is straight forward to check that complex conjugation defines an involution. In fact,
to a great extent complex conjugation inspired our definition of the involution. Hence,
Cb pXq is a Banach ‹-algebra. (Here we do not assume the involution to be continuous.
The continuity will follow, after we prove the C ‹ -property.) For the C ‹ -property we
observe for f P Cb pXq: |f pxqf¯pxq| “ |f pxq||f¯pxq| “ |f pxq||f pxq| “ |f pxq|2 , at any x P X.
Hence,
˘2
}f f¯} “ sup |f pxqf¯pxq| “ sup |f pxq|2 “ sup |f pxq| “ }f }2 . (11)
`
xPX xPX xPX
8
Comment: Indeed, by the Gelfand transform (that we meet in a few lectures) we
will see that every unital commutative C ‹ -algebra is isometrically isomorphic to a
C ‹ -algebra of continuous functions on some compact Hausdorff space.
Example 1.10. Consider our previous example: the algebra Mpn, Cq of n ˆ n matrices
with complex coefficients. Equipped with the operator norm and adjoint as an involution,
Mpn, Cq is a unital C ‹ -algebra.
Example 1.11. Let H be a separable Hilbert space. Denote by BpHq the space of linear
bounded operators on H. Then with the operator product, the operator norm, and the
adjoint as involution, the space BpHq is a unital C ‹ -algebra.
Proof. We only show the C ‹ -property as the rest is trivial. Let A be in BpHq then
}A}2 “ sup }Ax} “ sup xAx, Axy “ sup xx, A: Axy “ }A: A}. (12)
xPH; }x}“1 xPH; }x}“1 xPH; }x}“1
9
Take λ P Imgpf q, so for some x P X we have pf ´λ1qpxq “ 0. Since f ´λ1 is continuous
there is a sequence pxn q in X that converges to x and satisfies pf ´ λ1qpxn q ă 1{n. Let g
be the two-sided inverse such that pf ´ λ1qg “ gpf ´ λ1q “ 1. Then
1
(13)
` ˘
1 “ | pf ´ λ1qg pxn q| “ |gpxn q||pf ´ λ1qpxn q| ď |gpxq| ,
n
for every xn . Thus, |gpxn q| ě n, and g is not continuous at x. Consequently g R Cb pXq
and Imgpf q Ă spA pf q. Together with the opposite inclusion we get spA pf q “ Imgpf q.
Warning. For a Banach algebra the spectrum of an element does indeed depend
on the algebra. Therefore, we use the subscript to specify the ambient algebra. The
following example shows that changing the algebra can change the spectrum.
Example 1.15. Let H be a separable Hilbert space with the basis tej u for j P Z. Let BpHq
be the Banach algebra of linear bounded operators on H. (Above we have shown that
BpHq is a C ‹ -algebra, so in particular it is a Banach algebra). Consider the left shift L in
BpHq defined on the basis of H by Lpej q “ ej`1 and extended by linearity and continuity
to the whole H. The adjoint of L is the operator L: pej q “ ej´1 . One can easily verify that
LL: “ L: L “ 1. Thus, L: is the two-sided inverse of L. In other words, 0 R spBpHq pLq.
Now consider a Banach algebra B generated by 1 and L. It is a subalgebra of BpHq.
Is it still true that 0 is not in the spectrum of L relative to B? No. To see this consider
a closed subspace S of H spanned by tej : j “ 1, 2, . . . u. Then L maps this subspace
into itself, and thus, each polynomial of L leaves this subspace invariant. It follows that
every element in B (the norm closure of such polynomials) maps S into itself. Assume
that L̃ is a two-sided inverse of L in B. Then L̃ is also in BpHq since B Ă BpHq. By the
uniqueness of the two-sided inverse we thus have L: “ L̃. However, L: does not preserve
S as it maps the element e1 to e0 R S. Thus, L: is not in the algebra B. We can thus
conclude that L does not have a two-sided inverse in B and 0 P spB pLq.
The above example shows that the spectrum of an element depends on the ambient
Banach algebra. One of the main results that we will obtain in the first part of the lecture
will be that for C ‹ -algebras this cannot happen.
Lemma 1.16. For a Banach ‹-algebra A, let A be in A and let spA pAq be its spectrum.
Then the spectrum of A‹ is
(´ ¯
spA pA‹ q “ λ̄ : λ P spA pAq “ spA pAq . (14)
10
Proof. If λ R spA pAq then pA ´ λIq´1 exists in A. Hence
˘´1 ˘‹
(15)
` `
pA ´ λIq‹ “ pA‹ ´ λ̄Iq´1 “ pA ´ λIq´1 ,
˘´1
and pA ´ λIq‹ exists. Thus, λ̄ R spA pA‹ q. Interchanging A and A‹ in the above
argument yields the desired conclusion.
Lemma 1.17. For a Banach algebra A, let A be an invertible element in A. Then the
spectrum of A´1 is
´ ¯
´1 ´1
spA pA q “ tλ : λ P spA pAqu “ spA pAq ´1
(16)
Proof. If A is invertible then 0 R spA pAq. Thus, if λ is in spA pAq then λ ‰ 0 and
Since }A} ă 1 the right-hand side of the above estimate goes to zero. Thus, nk“0 Ak is a
ř
11
nÑ8
By continuity of the product we have }An`1 } ď }A}n`1 ÝÑ 0. Thus An`1 goes to zero
as n goes to infinity. Thence the right-hand side of (22) converges to I and we obtain
(again by continuity of multiplication)
on N is continuous.
The following definition will be convenient for the proof of the proposition.
Definition 1.20. Let A be an element of a Banach algebra A. We define the left multi-
plication by A as the mapping LA : A Ñ A such that LA pBq “ AB. In the similar way we
define the right multiplication by A as the mapping RA : A Ñ A such that RA pBq “ BA.
Info: From the (joint) continuity of the product in a Banach algebra it follows directly
that LA and RA are continuous for every A in A.
Proof of proposition 1.19. Even though the proof is not complicated, it is still helpful to
understand the general idea of the argument.
Idea: To show that N is an open set we need to show that for every element in
N there exists an open set containing this element and that is contained in N . We
know that an open ball of radius 1 around the identity I is in N . The idea is to shift
this ball appropriately to prove the proposition. The multiplication operators that
we just defined provide the “shifting”. We now formalize the idea.
12
LA
A A
N N
A A
Now we prove that the mapping inv is continuous on N . Assume that B P A is such
that }I ´ B} ă 1, then by the Neumann series
8
ÿ ˘´1
(25)
`
pI ´ Bqk “ I ´ pI ´ Bq “ B ´1 .
k“0
Thus inv is continuous at the identity. To show that it is continuous at every element
observe that for any B P N the following identity holds
Thus for each fixed element A P N choosing the right-hand side of the above equality
to be RA´1 ˝ inv ˝LA´1 shows that the map inv is a combination of continuous mappings
L inv R
A´1 . Thus for each A P N the mapping inv is continuous.
A ´1A ´1
A Ñ IÑI Ñ
We will now show that every element in a Banach algebra has a spectrum. The
proof uses some known, but not entirely trivial statements from the complex and func-
tional analysis. Since, these topics are themselves very extensive, we only recall the basic
needed statements before continuing with the theorem.
13
Reminder: Holomorphic function. A complex valued function f is called holo-
morphic at z0 P C if on some neighborhood of z0 the limit
f pzq ´ f pz0 q
lim (29)
zÑz0 z ´ z0
exists. A function that is holomorphic at every point of some open set O Ă C is said
to be holomorphic on O. If f pzq is holomorphic on an entire complex plain we call
it an entire function. By Liouville’s theorem, every bounded entire function must be
constant. In particular, if an entire function vanishes at infinite it must be zero.
The corollaries of this very general theorem are often more useful than the theo-
rem itself. For example the next corollary assures that we can always extend a linear
functional.
Another useful corollary is the one we use in the proof of Theorem 1.21.
Equipped with all that fundamental knowledge. We can now prove that the spectrum
of an element in a Banach algebra is never empty.
Theorem 1.21. If A is an element of a Banach algebra A then spA pAq is a non-empty closed
subset of the closed disc in C with center 0 and radius }A}.
14
Proof. Let λ P C be such that |λ| ą }A} then
´1 ¯
A ´ λI “ λ A´I . (32)
λ
Since }Aλ´1 } ă 1 the element λ1 A ´ I is invertible (by the Neumann series). Thus, A ´ λI
has the inverse λ1 p λ1 A ´ Iq´1 . It follows that spA pAq Ă D}A} p0q (a disc with center 0 and
radius }A}).
Now we need to show that spA pAq is not empty. Let λ R spA pAq. Then by the proposi-
tion 1.19, the element A´λI is invertible in a small neighborhood around λ. In particular
the set Cz spA pAq is open.
Reminder: Resolvent equation. Recall the first resolvent equation [3, thm. VI.5]
The above theorem states that the spectrum of an element in the Banach algebra is
bounded by its norm. However, the norm is just an upper bound not the smallest upper
bound. To study how different or similar is the norm and the smallest upper bound of
the spectrum, we give the latter quantity a proper name first.
Definition 1.22 (Spectral radius). The spectral radius rA pAq of an element A in a Banach
algebra A is
15
C
}A}
spA pAq
rA pAq
Info: The spectral radius is the radius of the smallest disc in C with center 0 contain-
ing spA pAq. Also from the preceding theorem, }A} is the radius of the largest disc in
C with center 0 that contains spA pAq. Hence we always have rA pAq ď }A}.
16
Hence, if q “ 2m for some m “ 1, 2, . . . then by induction we get
}H q } “ }H}q . (38)
So the assertion holds for self-adjoint elements. Now, let A be normal and H “ A‹ A then
With the following justifications for the estimates: (1) is the C ‹ -property; (2) follows
from what we have just shown above; (3) follows from the fact that for normal elements
}pA‹ Aqn } ď }pA‹ qn }}An } and thus lim }pA‹ Aqn }1{n ď lim }pA‹ qn }1{n }An }1{n ; (4) follows
from the fact that spA pA‹ q “ spA pAq; (5) follows since }A} defines the largest radius of
a disc that contains the spectrum.
Since the left and the right hand side of the above estimate are equal all the "less or
equal" signs in the estimate have to be actual equalities. Hence rA pAq “ }A}.
So combining (41) and (42), we get a2 ` p2n ` 1qb2 ď }A}2 for every n P t0, 1, 2, . . . u.
Since A is, however, bounded b must be zero.
17
Algebra vs. bounded operators on a Hilbert space. For bounded operators on
a Hilbert space we can use the inner product to prove that a self-adjoint operator
on a Hilbert space has a real spectrum. This makes the proof simpler (but also less
general). Consider a self-adjoint bounded linear operator A on a Hilbert space H.
Let λ P C be such that | Impλq| ą 0. Then for every x P H we have by the Cauchy-
Schwarz inequality,
18
Hence, we get
` ˘ ` ˘
ppAq ´ ppλqI “ pA ´ λIq an pAn´1 ` λAn´1 ` ¨ ¨ ¨ ` λn´1 Iq ` ¨ ¨ ¨ ` a1 “ ¨ ¨ ¨ pA ´ λIq,
and ppAq ´ ppλqI does not have a two-sided inverse (otherwise A ´ λI would have a
two-sided inverse). It follows that ppλq P spA pppAqq, and thus ppspA pAqq Ă spA pppAqq.
For the opposite inclusion consider that if γ is in spA pppAqq and λ1 , . . . , λn are the n
roots of ppxq ´ γ then
and at least one of A ´ λj I is not invertible. Hence, λj P spA pAq for at least one of
λj ’s and by previous discussion γ “ ppλj q “ spA pppAqq (the first equality follows since
ppλj q ´ γ “ 0 by definition of roots). Hence, spA pppAqq Ă ppspA pAqq.
To unclutter the notation in the proofs we will use the following common notations.
We summarize them in the following paragraph.
• by CpspA pAqq we will denote the C ‹ -algebra of continuous functions on spA pAq;
It is time to introduce the function calculus for self-adjoint operators. Almost the same
theorem, with minor modifications, holds also for normal elements in a C ‹ -algebra (see
the Remark at the end of the section). However, we will prove only the following result,
as it does not require any additional technicalities that are needed for normal elements.
Before the proof, we recall the Weierstrass approximation theorem for convenience.
19
iii ) ϕpf q is the limit of a sequence of polynomials in I and A;
iv ) if B is in A that commutes with A, then it also commutes with ϕpf q.
By Proposition 1.24 ii ), the spectrum of A is a compact subset of the real line. Then we
can write p restricted to this subset using the identity mapping in CpspA pAqq such that
n
ÿ
p“ ak id k pwith id 0 “ 1q. (48)
k“0
Define ϕ0 : P l Ñ A by
n
ÿ n
ÿ
ϕ0 ppq “ ak Ak “ ppAq and ϕ0 ppq‹ “ āk Ak “ ppAq‹ . (49)
k“0 k“0
Then, ϕ0 is linear and ϕ0 pid k q ÞÑ Ak for k “ N. Also, ϕ0 ppq and ϕ0 ppq‹ commute, implying
that ϕ0 ppq is normal. Thus, Proposition 1.24 i ) applies and together with Proposition 1.25
we have
}ϕ0 ppq} “ rA pppAqq “ supt|λ| : λ P spA pppAqqu “ supt|ppλq| : λ P spA pAqu “ }p}, (50)
where the norm on the left-hand side is the C ‹ -norm on A and the norm on the right-
hand side is the supremum norm on CpspA pAqq. Thus, ϕ0 is an isometry. If p and q are
two polynomials in P l, identically equal on spA pAq, then
and then ϕ0 ppq “ ϕ0 pqq. Thus, the mapping ϕ0 is well defined (unambiguous).
By the Weierstrass approximation theorem, P l Ă CpspA pAqq is everywhere dense in
CpspA pAqq. Since A is complete, ϕ0 extends to the mapping ϕ defined on CpspA pAqq. This
extension remains isometric; since if f P CpspA pAqq and tpn u is a sequence of polynomials
that approximates f then,
20
therefore it contains ApAq. Hence ApAq is the smallest C ‹ -subalgebra that contains I and
A. (This proves ii ) and iii ).)
By definition, ϕ0 is linear, multiplicative (i.e. ϕppqq “ ϕppqϕpqq for p, q P P l), preserves
the star (i.e. ϕpp̄q “ ϕppq‹ ), and carries the unite of P l onto that of A. These properties
extend by continuity to ϕ and thus, ϕ is a ‹-homomorphism. Since it is isometric it is
one-to-one and thus, a ‹-isomorphism. (This proves the isometry assertion.)
If φ is another such ‹-isomorphism that maps id to A then φ and ϕ coincide on the
dense space P l; thus they are equal. (This proves uniqueness of the asserted mapping.)
Since we can approximate every element in ApAq by polynomials in A, it follows at
once that ϕpf q is a normal element, and that it is self-adjoint if and only if f is real
valued. If B P A commutes with A then it commutes with all polynomials in A, and thus
it commutes with every element in ApAq. This concludes the proof.
The mapping ϕ is called the function calculus for the self-adjoint element A. To
simplify the notation, it is often common to write f pAq instead of ϕpf q. Henceforth, we
will use this convention.
We have already seen that a spectrum of a Banach algebra depends on the ambient
Banach algebra. For C ‹ -algebras this is not the case. Here, the spectrum of an element
is independent of the C ‹ -algebra in which we consider this element to be. Stated sloppy,
C ‹ -algebras are always large enough to contain an inverse of an element if it exists. With
the aid of the function calculus we can now prove this statement.
Proof. We have already discussed that spA pBq Ď spB pBq. Thus we need only to prove
the opposite. For this we will show that if A P B has an inverse in A then this inverse is
also in B.
We consider first a self-adjoint element A. Since 0 R spA pAq, the function f ptq “ t´1
is continuous on spA pAq. Thus, by the function calculus relative to the algebra A we get
an element f pAq in A. From Theorem 1.26 ii ) this element is contained in B. Since
tf ptq “ 1 for each t in spA pAq, it follows by multiplicativity of the function calculus that
21
The above proposition shows that the spectrum of an element in a C ‹ -algebra is inde-
pendent of the C ‹ -algebra this element is in. Thus, from now on, we can (and will) omit
the subscript referring to the ambient C ‹ -algebra and write sppAq instead of spA pAq.
In the proof of Proposition 1.27 we use the phrase “... by the function calculus relative
to the algebra A ...”. Implying, that the function calculus may depend on the original al-
gebra. Nevertheless, using Proposition 1.27 we can see that this cannot happen. Namely,
if B is a C ‹ -subalgebra of A, and A is a self-adjoint element in B, then let ϕ be the spec-
tral calculus sppAq Ñ A and ψ be the spectral calculus sppAq Ñ B. However, since the
spectrum of A relative to either algebra is the same, we can consider ψ as a spectral calcu-
lus into the larger algebra A. By uniqueness of the spectral calculus, ψ must then coincide
with ϕ. Thus, we do not need to worry about the ambient C ‹ -algebra when talking about
the spectral calculus. In particular, we can always choose a sufficiently small C ‹ -algebra,
ApAq “ tf pAq : f P CpsppAqqu such that the spectral calculus, CpsppAqq Ñ ApAq, is onto.
Then, we can state the following about the spectrum of f pAq.
Proof. The function calculus f Ñ f pAq is a ‹-isomorphism from CpsppAqq onto the C ‹ -
subalgebra ApAq of A, and ‹-isomorphisms preserve the spectrum.
Another useful consequence of the function calculus is the definition of a square root.
Since, f is real valued, the element H is self-adjoint by 1.26 i ). Since f is positive the
spectrum of H is positive by 1.28. If B P A commutes with A then it commutes with H
by 1.26 iv ).
To show uniqueness, suppose that K is any element in A with positive spectrum and
?
such that K 2 “ A. Let tpn u be a sequence of polynomials that converge to f “ id. Set
qn ptq “ pn pt2 q. Then we have,
22
uniformly for t P sppKq. Hence, by the function calculus for K we get
There are more useful corollaries that we can derive from the function calculus. For
now, however, we close this chapter and proceed to the next ingredient that will be
necessary to understand the construction of quantum mechanics: the GNS construction.
Info: The function calculus can be extended to normal elements. The methods to
prove it, however, are a bit different. We provide the full theorem without proof.
Theorem 1.30 (Function calculus for normal elements). Let A be a normal element
of a C ‹ -algebra A. There is a unique isometric ‹-isomorphism ϕ from CpspA pAqq into
A such that ϕpid q “ A. Moreover, for f P CpspA pAqq,
i ) ϕpf q is a normal element of A;
ii ) the set tϕpf q : f P CpspA pAqqu is a commutative C ‹ -algebra and it is the smallest
C ‹ -subalgebra of A that contains A;
iii ) ϕpf q is the limit of a sequence of polynomials in I, A, and A‹ ;
iv ) moreover, the element A is
23
1.3 The GNS construction
In this section we will see that any abstract C ‹ -algebra can be “represented” as an al-
gebra of bounded linear operators on some appropriate Hilbert space. The construction
used to show this was developed by Gelfand, Naimark, and Segal, hence the name GNS-
construction. This construction is vital for our purpose: to use C ‹ -algebras in applications
to quantum mechanics. To discuss the GNS construction we need additional tools — the
order structure on the C ‹ -algebra, the space of states, and a brief discussion of represen-
tations of a C ‹ -algebras. We begin with the order structure.
Reminder: partial order. A partial order on a set Ω is a relation over the set Ω such
that if a, b, c P Ω:
i ) a ď a, that is the relation is reflexive;
ii ) if a ď b and b ď a then a “ b, that is the relation is anti-symmetric;
iii ) if a ď b and b ď c then a ď c, that is the relation is transitive.
It is easy to check that C is indeed a cone, and thus this provides a partial ordering for
the space R2 . However, this cone (and thus the partial ordering) depends on the basis
and is not preserved under rotations.
24
aA
A`B
B
B´A
A
B A
(a) A cone in a vector space. (b) Partial order induced by the cone.
Since a (positive) cone defines a set of positive elements, it stands to reason that a
set of positive elements — defined in some meaningful way — will yield a cone. As is
common, we can borrow the intuition of positive elements from bounded linear operators
to define positive elements in a C ‹ -algebra.
Example 1.35. If B is a bounded linear operator on some Hilbert space H then we call
it positive, if xx, Bxy ě 0 for every x P H. We have seen in our tutorial class, that B is
positive if and only if it is self-adjoint and its spectrum is positive. Hence, the notion of
positivity for elements in the C ‹ -algebra BpHq is consistent with the notion of positivity
for bounded linear operators.
Even though, positivity is a statement about “signs”, it has a very useful characteri-
zation in terms of the norm. We state this characterization in the following lemma.
}A ´ aI} ď a, (60)
25
Proof. Fix a ě }A}, then sppAq Ď r´a, as, and
Where the first equality holds by 1.24 i ) and the second equality holds by 1.25. Hence
}A ´ aI} ď a if and only if sppAq Ď R` .
Now, we can show that the set of positive elements in a C ‹ -algebra defines a closed
positive cone.
(64)
› › › ›
›A ´ }A}I › ď }A} ›B ´ }B}I › ď }B}.
Therefore,
(65)
› ›
›A ` B ´ p}A} ` }B}qI › ď }A} ` }B}.
26
Condition, ii ), iii ), and iv ) state that A` is a (positive) cone. Hence, we can use it
to define partial ordering on the C ‹ -algebra A. Using this ordering, we can say that an
element A is positive if A ě 0. Also notice, that in one of our exercises we have shown
that }A}I ´ A is always positive, if A is self-adjoint. Hence, for a unital C ‹ -algebra the
unit of the algebra is also the order unit for this partial ordering,
So far, we have discussed the elements of the algebra. However, a great deal of structure
on the C ‹ -algebra comes from its dual space: the space of continuous linear functionals
on the algebra. We have encountered the power of linear functionals in the Theorem
1.21. Now we approach the topic of linear functionals more systematically. With the aid
of the positive cone on the C ‹ -algebra we can define a cone on its dual space.
Definition 1.38 (Convex set.). A set Y in a vector space V is called convex if x and
y in Y imply that the line tby1 ` p1 ´ bqy2 : 0 ď b ď 1u is also in Y .
Definition 1.39 (Locally convex space.). A locally convex space is a topological vector
space, in which the topology has a base consisting of convex sets. That means: Let
V be a topological vector space with topology τ . The basis of τ is the family β Ă τ
such that every open set O P τ is of the form O “ Yα Bα for some family tBα u Ă β.
If each set in β is convex, then V is a locally convex space.
After this short reminder we are equipped with everything to dive into the discussion
of linear functional on a C ‹ -algebra. We begin with some general definitions.
Definition 1.42. Let V be a partially ordered vector space with order unit I. Then,
i ) a linear functional ρ on V is called positive if ρpvq ě 0 for every v ě 0 with v P V ;
ii ) a positive linear functional ρ on V is called a state if ρpIq “ 1;
iii ) the state space SpV q is the set of all states on V .
27
If ρ is a linear functional on a C ‹ -algebra A, we can define another linear functional
ρ‹ , by the equation ρ‹ pAq “ ρpA‹ q for every A P A.
Definition 1.43. Let A be a C ‹ -algebra and ρ a linear functional on it. If ρpAq “ ρpA‹ q
for every A P A, we call ρ hermitian.
Recall that every element of a C ‹ -algebra uniquely decomposes into its real and imag-
inary part. Using this decomposition we can rephrase the condition for a state to be
hermitian only in terms of self-adjoint operators: let A be an element in a C ‹ -algebra A
with the real part H and imaginary part K, then for a linear functional ρ on A we have,
ρpAq “ ρpH ` ıKq “ ρpHq ` ıρpKq, ρpA‹ q “ ρpH ´ ıKq “ ρpHq ` ıρpKq. (67)
It follows that ρ is hermitian if and only if ρpHq “ ρpHq for every self-adjoint H P A.
Using this characterization of hermitian functionals we can deduce that a positive
linear functional is always hermitian: Let A be self-adjoint in A and ρ a positive linear
functional on A. Since }A}I ˘ A ě 0 (as we already have seen) it follows from positivity
of ρ that ρp}A}I ˘ Aq ě 0. In particular, ρp}A} ˘ Aq is real, and from
1´ ` ˘¯
(68)
˘ `
ρpAq “ ρ }A}I ` A ´ ρ }A}I ´ A ,
2
it follows that ρpAq itself is real.
Additionally, we can characterize positive linear functionals by their norm as follows.
Proof. For the “only if ” direction, assume that ρ is positive. Let A be in A, choose a P C
such that |a| “ 1 and aρpAq ě 0, and let H be the real part of aA. Since }H} ď }A} we
have H ď }H}I ď }A}I and thus
28
So by the spectral radius formula we have }I ´ sA} “ rpI ´ sAq ď 1. From the continuity
of ρ we then have
where the last inequality follows from the continuity of ρ and (73). Since A is bounded,
b has to vanish.
With this understanding we can summarize the similarities and differences between
positive elements in the algebra and its dual.
An element A in A is positive if An element ρ in A1 is positive if
i ) it is self-adjoint, i ) it is hermitian,
ii ) sppAq Ă R` , ii ) ρpAq ě 0 whenever A P A` ,
iii ) }A ´ aI} ď a. iii ) ρpIq “ }ρ} ă 8.
Before we proceed with the analysis of states, we recall some ideas regarding the
topologies on vector spaces.
Reminder: weak‹ -topology. Let V be a linear space and let F be a family of linear
functionals on V . Assume that F separates points of V in the sense that if x P V is
non-zero then there is a ρ P F such that ρpxq ‰ 0. We define the topology σpV, Fq
as the coarsest (weakest, one with the fewer open sets) topology on V relative to
which each element of F is a continuous mapping from V into C.
If V is a locally convex topological vector space, we denote its continuous (topo-
logical dual) as V 1 . If x is in V then the equation
defines a linear functional x̂ on the dual V 1 . The set V̂ “ tx̂ : x P V } is a linear sub-
space of V 2 — the algebraic dual space of V 1 (the space of not necessarily continuous
functionals on V 1 ). The set V̂ separates points of V 1 since, if ρ P V 1 and x̂pρq “ 0 for
all x P V then ρpxq “ 0 for all x P V and thus ρ “ 0. The weak‹ -topology on V 1 is the
topology σpV 1 , V̂ q. We often denote it simply by σpV 1 , V q. Hence, the weak‹ -topology
29
is the coarsest topology on V 1 for which each functional x̂ is continuous.
A bounded closed set in an infinite dimensional space is not always compact. How-
ever, in the weak‹ -topology this is indeed true (see the useful Lemma below). This makes
the weak‹ -topology an important tool in analysis.
For A P A let  be the dual element of A1 such that ρpAq “ Âpρq for every ρ P A1 .
Since each  is by definition continuous in the weak‹ -topology, the set CA “ ´1 r0, 8q
` ˘
— the pre-image of r0, 8q under  — is closed, and so is the set C0 “ Iˆ´1 pt1uq. Then
SpAq “ XAPA` CA XC0 is an intersection of weak‹ -closed sets and thus is weak‹ -closed.
` ˘
In summary: SpAq is a bounded, weak‹ -closed, convex subset of A1 , and thus from the
“useful lemma” above, it is weak‹ -compact. By the Krein-Milman theorem SpAq is the
closed convex hull of its extreme points.
Definition 1.45. An extreme point of SpAq is called a pure state of A. The set of all pure
states of A will be PpAq. The weak‹ -closure of PpAq is called the pure state space of A.
Visually we can picture the state space as in Figure 4. In general, the set of pure states
is not weak‹ -closed, and PpAq can contain states which are not pure. For commutative
C ‹ -algebras, however, this does not happen. So if A is commutative then the space of all
pure states is already weak‹ -closed. This observation relates to a deep result attributed
to Gelfand — sometimes called the Gelfand representation of a C ‹ -algebra. We state the
result without proof.
30
SpAq
mixed states
PpAq
Figure 4: Intuition for the space of states SpAq. The pure states lie on the black solid
boundary line. The line, however, is not closed; it may miss some points, which are
indicated in red. The weak‹ -closure of that line is the whole boundary and it is the space
of pure states PpAq. In particular, the space of pure states may contain elements which
are not pure states (the red points). The interior of the set is sometimes called the space
of mixed states.
Up to now we discussed abstract C ‹ -algebras. Now, we will see that every C ‹ -algebra
can be represented as a (sub)algebra of bounded linear operators on some Hilbert space.
From the physicist point of view, it is this realization that allows us to formulate quantum
and classical mechanics in terms of algebras. After all, as a physicist we understand
operators in terms of how they act, not in terms of to what space they belong.
31
The Gelfand-Naimark theorem assures that every C ‹ -algebra has a faithful represen-
tation on a Hilbert space. However, we only will discus the construction that provides a
(possibly not faithful) representation.
The following terminology is common when discussing representations. Let ϕ be a
representation of a C ‹ -algebra A on a Hilbert space H. Recall from our assignment 4 i)
that every ‹-homomorphism between C ‹ -algebras is continuous, and so ϕ is continuous,
i.e. }ϕpAq}H ď }A}A for every A P A. Also recall, that a ‹-isomorphism (an injective
‹-homomorphism) is isometric, and so if ϕ is faithful }ϕpAq}H “ }A}A . If you recall the
definition of ideals from assignment sheet 5, you realize that the set A P A : ϕpAq “ 0
(
is a closed two-sided ideal in A. We call this set the kernel of ϕ. If there is a vector x P H
for which the linear subspace
(77)
(
ϕpAqx “ ϕpAqx : A P A
Example 1.49. Let A be a subalgebra of bounded linear operators BpHq on the Hilbert
space H. Then the inclusion mapping ϕpAq “ A for every A P A is a representation. This
representation is faithful, since if A P A is not zero, then it is also non-zero in BpHq.
Let H be a Hilbert space K be a closed subspace of H, that is invariant under each
operator from A, i.e. the set Ax : A P A, x P K is a subset of K. Let A|K be the
(
restriction of the operator A to the subspace K. It can be checked that this restriction
is a ‹-homomorphism and thus A Ñ A|K : A Ñ BpKq is a representation of A on the
Hilbert space K. In the same way one can see that the orthogonal complement K K is
also invariant under A and thus the restriction A Ñ A|K K is a representation of A on K K .
If the inclusion representation A P A Ñ A P BpHq is cyclic then also the restriction
representations are cyclic as well. To see this let P be the orthogonal projection from H
onto K and x P H be the cyclic vector x P H for the inclusion representation. If P x “ 0
then x is in K K . This, violates the fact that the inclusion representation is cyclic since
K K is invariant under the action of A. Thus P x ‰ 0 and P x P K is a cyclic vector for the
representation A Ñ A|K , and pI ´P qx is a cyclic vector for the representation A Ñ A|K K .
32
Reminder: Inner product and the Cauchy-Schwarz inequality. In the following
we use the convention that an inner product on a complex vector space V may not
be definite. That is, we say that a mapping pv, wq Ñ v, w from V ˆ V to C is an
@ D
ii ) hermitian: v, w “ w, v ;
@ D @ D
iii ) sesquilinear: v, αw ` βz “ α A, B ` β A, C ;
@ D @ D @ D
@ D
Proposition 1.50. Let V be a complex vector space. Suppose that , is an inner
product on V . Then,
@ D @ D@ D
i ) the Cauchy-Schwarz inequality, | v, w |2 ď v, v w, w , holds for v, w P V ;
@ D
ii ) the set L “ tv P V : v, v “ 0u is a linear subspace of V ;
iii ) the equation
(78)
@ D @ D ` ˘
v ` L, w ` L q “ v, w v, w P V
@ D
defines a definite inner product , q on the quotient space V {L.
Proof. For the proof of the above Proposition consult e.g. [1, prop. 2.1.1]
(
Proposition 1.51. Let A be a C ‹ -algebra, ρ a state on A, and Lρ “ A P A : ρpA‹ Aq “ 0 .
The state ρ provides a definite inner product on the quotient space A{Lρ by the relation
(79)
@ D ` ˘
A ` Lρ , B ` Lρ “ ρpA‹ Bq A, B P A ,
product on A{Lρ .
With A P Lρ and B P A we have by the Cauchy-Schwarz inequality
(81)
` ˘ ` ˘
ρ pBAq‹ BA “ ρ A‹ pB ‹ BAq “ 0,
33
and thus BA is in Lρ whenever A P Lρ and B P A. Hence, Lρ is a left ideal of A, which
is closed since ρ and the mapping A Ñ A‹ A are continuous.
We will call the space Lρ the left kernel of the state ρ. We can now summarize the
idea of the GNS-construction in a loosely way before proving it rigorously. The preceding
Proposition provides us with means to define a (definite) inner product on A. With the
norm induced by this inner product, the vector space A may not be closed, but after
closing it we obtain a Hilbert space. Moreover, since A is an algebra we can define an
action of A P A on the vector space A by setting B Ñ AB : A Ñ A. After proving that
this mapping is continuous we can extend it to the whole H. This will give us the desired
representation. As always, however, the devil lies in the complement of a dense set...
Theorem 1.52 (GNS representation). For every state ρ on a C ‹ -algebra A there is a cyclic
representation πρ of A on a Hilbert space Hρ and a unit cyclic vector xρ for πρ such that
(82)
@ D ` ˘
ρpAq “ xρ , πρ pAqxρ APA .
If ϕ is a cyclic representation of A on a Hilbert space H with the unit cyclic vector x for π
such that
(83)
@ D ` ˘
ρpAq “ x, ϕpAqx APA ,
(84)
` ˘
x “ U xρ and ϕpAq “ U πρ pAqU ˚ APA .
Proof. Let Lρ be the left kernel of ρ. The quotient linear space A{Lρ is a pre-Hilbert space
relative to the definite inner product defined by
(85)
@ D ` ˘
A ` Lρ , B ` Lρ “ ρpA‹ Bq A, B P A .
Define Hρ as the completion of A{Lρ in the norm induced by the above scalar product.
To show that for A P A the mapping B ` Lρ Ñ AB ` Lρ defines an unambiguous
operator on Hρ , we need to show that it does not depend on the “representative” B of
the set B ` Lρ . For this assume B, C P A such that B ` Lρ “ C ` Lρ . Then B ´ C is
in Lρ and since Lρ is a left ideal of A we have for every A P A that AB ´ AC P Lρ and
consequently AB ` Lρ “ AC ` Lρ . Hence, the equation πpAqpB ` Lρ q “ AB ` Lρ defines
a linear operator πpAq acting on the pre-Hilbert space A{Lρ .
Next, we want to show that π is continuous. For this recall that
34
˘1{2
implies that the positive square root K “ }A}2 I ´ A‹ A exists, and B ‹ p}A}2 I ´
`
Thus πpAq is bounded, with }πpAq} ď }A}; and πpAq extends by continuity to a bounded
linear operator πρ pAq acting on Hρ . It remains to show that πρ is a ‹-homomorphism and
that it is cyclic.
With A, B, C P A and α, β P C we get:
linearity,
multiplicativity,
` ˘ ` ˘ ` ˘
πρ pABq C ` Lρ “ ABC ` Lρ “ πρ pAq BC ` Lρ “ πρ pAqπρ pBq C ` Lρ ;
star preservation,
@ D @ D
B ` Lρ , πρ pAqpC ` Lρ q “ B ` Lρ , AC ` Lρ
` ˘ ` ˘
“ ρ B ‹ pACq “ ρ pA‹ Bq‹ C
@ D @ D
“ A‹ B ` Lρ , C ` Lρ “ πρ pA‹ qpB ` Lρ q, C ` Lρ .
Since A{Lρ is everywhere dense in Hρ these properties extend to the whole Hρ and thus
πρ is a representation of A on Hρ .
To see that πρ is cyclic let xρ be the vector I ` Lρ in A{Lρ , then
(87)
` ˘
πρ pAqxρ “ πρ pAqpI ` Lρ q “ A ` Lρ APA ,
35
U0 : πρ pAqxρ Ñ ϕpAqx by the equation U0 πρ pAqxρ “ ϕpAqx for each A P A. It follows
that U0 is a norm-preserving linear operator from πρ pAqxρ onto ϕpAqx. Since the closure
of πρ pAqxρ is the entire Hρ and the closure of ϕpAqx is the entire H, it follows that U0
extends by continuity to an isomorphism U from Hρ onto H, and
With A, B P A,
The method used in the proof of the above Theorem is called the GNS (Gelfand-
Naimark-Segal) construction. To avoid bloated notation, we will often write pπ, H, xq to
refer to a cyclic representation π of a C ‹ -algebra A on the Hilbert space H with a cyclic
vector x for π. When its convenient we will also call the triple pπ, H, xq itself a cyclic
representation of A. If ρ is a state on A, and pπρ , Hρ , xρ q is the representation set out in
the Theorem 1.52, we will call it the GNS representation.
For a given C ‹ -algebra A, let π be a representation of A on the Hilbert space H
and ϕ be another representation of A on the Hilbert space K. We will call the two
representations (unitarily) equivalent if there exists an isomorphism U : H Ñ K such
that ϕpAq “ U πpAqU ‹ for each A P A. In particular, the GNS Theorem states that a
GNS representation πω of A and the cyclic state xω is uniquely determined (up to unitary
equivalence) by the condition ρp¨q “ xω , πω p¨qxω .
@ D
36
Example 1.53. Consider a C ‹ -algebra A and a state ρ on A.
i ) The GNS representation for the state ρ is the ‹-homomorphism πρ : A Ñ BpHρ q,
with the Hilbert space Hρ and a cyclic (GNS) vector xρ ;
ii ) Let y be a unit vector in Hρ . Then we can use y instead of the cyclic vector for πρ . If
y is cyclic for πρ then pπρ , Hρ , yq is a new cyclic representation which is equivalent
to the GNS representation πρ . In fact, the ‹-homomorphism and the Hilbert spaces
are both the same, and thus U “ I is the desired isomorphism. (Notice that we do
not need that the cyclic vectors transform from one to the other).
iii ) For y from the previous example, set ωy pAq “ y, πω pAqy for every A P A. Then
@ D
SpAq H SpAq
BpH1 q
BpHn q
Figure 5: The state space SpAq yields different GNS representations. Some are equiva-
lent some are not. In particular, each (cyclic) vector from the GNS Hilbert space defines
a state. The GNS representations of these states are all equivalent to the original repre-
sentation. Also other states (not created from vectors from the Hilbert space) can yield
equivalent representations.
37
2 Operator algebras in physics
Section remarks: This section is based on two books. The part on classical me-
chanics as well as the formulation of the Dirac-von Neumann axioms is adopted from
[6]. The axioms for a general physical system are from [7]. Nevertheless, axiom 4 is
shortened, because otherwise we would need too much additional notation. Also the
final axiom from [7] is missing entirely, because by now we have a classification of
Jordan-Banach algebras that make the axiom (more or less) obsolete. The operational
motivation (viewing observables as measurement devices) is again from [6]. Further
reading: the book by Strocchi [6] is very commendable for a simple introduction to the
physical side of the discussion. Despite being a book for mathematicians, it focuses on
the general picture to phrase quantum mechanics in the algebraic language. It is easy
to read and is not too heavy on the mathematical details. As a complementary read
the book by Emch [7] is mathematically much more precise and helps to understand
how careful and rigorous the algebraic formulation is constructed. For the algebraic
formulation of statistical physics Ruelle [8] is a great reference. Alternatively, the two
volumes by Bratteli and Robinson [9, 10] give a self-consistent introduction to the sub-
ject. The first volume covers the C ‹ -algebra theory, the second volume discusses the
algebraic formulation of quantum mechanics and non-relativistic quantum field the-
ory. Contrary to Strocchi and Emch, however, these books focus almost exclusively on
the mathematics. The deep physical ideas that historically lied underneath the subject
are not a matter of discussion in these two books.
2.1.1 Kinematics
38
(generalized) positions and (generalized) momenta of a particle. The state of a mechan-
ical system is described by a set of canonical variables, tq, pu where q “ pq1 , . . . qn q P Rn
and p “ pp1 , . . . , pn q P Rn . The collection of all possible canonical variables for the system
is called the phase space Γ. For simplicity, we will assume in the following that the phase
space is compact, that is the range of possible positions and momenta of a particle are
bounded. Since the kinetic energy of a particle with the mass m is given by
p2
Ekin pq, pq “ , (90)
2m
this assumption means that the particle under consideration is confined to a bounded
region and has finite energy. Thus this assumption is reasonable.
Observables of the system are the physical quantities that result from the physical de-
grees of freedom. Thus in particular the position q and the momentum p are themselves
observables. But all its powers and sums are observables too, and so the observables
include the polynomials of the canonical variables q and p. Without loss of generality, we
can take the closure of this space in the uniform topology (supremum norm), to obtain
that all observables form the space of real-valued continuous function CpΓqR .
Equipped with a pointwise multiplication CpΓqR forms a real algebra. At the same
time, CpΓqR is the subspace of self-adjoint elements of the C ‹ -algebra CpΓq equipped
with the conjugation as the star operation and topologized with the supremum norm.
Extending CpΓqR to CpΓq is quite natural from the algebraic point of view. It is also not
harmful from the physics point of view since a C ‹ -algebra is uniquely determined by its
self-adjoint elements as we have seen. Thus, we see that the space of observables of a
classical system defines a C ‹ -algebra with identity I — the constant unit function on Γ.
Physical states define the physical system at hand. From our description, every state
is a point of the phase space tq, pu “ P P Γ, and it uniquely determines the value of
all the observables. Conversely, the value of all the observables uniquely determines the
state by the Urysohn’s and the Stone-Weierstrass theorem. This is sometimes called the
duality relation between the states and observables.
Identifying states of a classical system with points of a phase space is, however, an
idealization. It requires infinitely precise measurements to resolve single points of Γ. For
realistic measurements, we always have measurement errors and noise that corrupt the
measurement. The general procedure (and the one that is in fact used in physics) takes
into account these errors. To overcome the noise problem we typically repeat the exper-
iment several times and take the average of the measurement results. More precisely, let
us refer to the state of a classical system by ω. Then the value of an observable f P CpΓq
in the state ω is given by the following procedure: we take n replicas of ω and perform
measurements of f on each of the replicas. With this we obtain the measurement results
39
m1 pf q, m2 pf q, . . . , mn pf q; subsequently we compute the average
pωq pωq pωq
@ Dpωq 1 ` pωq
(91)
pωq ˘
f n “ m1 pf q ` m2 pf q ` ¨ ¨ ¨ ` mpωq
n pf q
n
It is one of the basic assumptions of experimental physics that for n Ñ 8 the above
quantity converges. This limit value is the expectation value of the observable f in the
state ω. It is this value that we assign to a measurement
@ Dpωq
ωpf q “ lim f n . (92)
nÑ8
Since the expectation values of all the observables is all there is that we can measure,
we can identify a state of a physical system with the set of expectation values of its
observables. Since ωpf q is operationally defined by the averaging process it follows that
such expectations are linear. That is, for f, g P CpΓq and α, β P C we have
(94)
` ˘
ωpf ‹ f q ě 0 f P CpΓq ,
(95)
@ D ` ˘
f, g “ ωpf ‹ gq f, g P CpΓq ,
(96)
@ D @ D@ D ` ˘
| f, g |2 ď f, f g, g f, g P CpΓq .
If ωpIq “ 0 then ωpf q “ 0 for every f P CpΓq, meaning that ω is a trivial functional.
Conversely, if ω is non-trivial then ωpf q ‰ 0 for at leas one f P CpΓq and thus ωpIq ‰ 0.
In this case, we can normalize ω to satisfy ωpIq “ 1. Thus we can identify physical states
of a classical system with (algebraic) states on the algebra of observables CpΓq.
40
Proof. For the proof consult e.g. [3, thm. IV. 14].
From the Riesz-Markov theorem it follows that a state ω on CpΓq defines a Radon
measure (regular Borel measure) µω on Γ such that
ż
(99)
` ˘
ωpf q “ f pq, pq dµω pq, pq f P CpΓq .
Γ
Since 1 “ ωpIq “ Γ dµω “ µpΓq we see that µ defines a probability measure on Γ. Thus
ş
If P “ tq0 , p0 u we have
ż
δP pf q “ f pq, pq dµδP pq, pq “ f pP q “ f pq0 , p0 q. (101)
Γ
Thus δP is an evaluation state, and from the Gelfand representation Theorem 1.46 and
the remark that follows it, we saw that such evaluation states define pure state of the
C ‹ -algebra CpΓq. Thus the standard description of classical mechanics, where points of
the phase space define the states, corresponds to pure states in the algebraic formalism.
To quantify the precision of measurements for an observable f P CpΓq in a state ω,
we use the variance
(102)
` ˘
p∆ω f q2 “ ω pf ´ ωpf q2 q .
The smaller the variance, the sharper is the measurement — the more “reliable” is the
mean value ωpf q. Pure states such as δP (P P Γ) minimize the variance, as for them we
have δP pf 2 q “ δP pf q2 “ f pP q2 and thus ∆δP f “ 0. Thus such states are called dispersion-
free, as they represent measurements of a classical system without any uncertainty.
Since algebraic states capture both, the pure and the mixed states, they provide a uni-
fied framework to define sharp measurements as well as statistical fluctuations. For that
reason we can use the same formalism to describe statistical mechanics. In fact, in sta-
tistical mechanics the description of the system with pure states is unfeasible because we
would need to know the exact positions and momenta for 1023 particles. Therefore, the
only reasonable description of such systems is using probability distributions for states
and random variables to describe observables. This is what we usually encounter in sta-
tistical approach to thermodynamics: the partition function there defines a probability
distribution, and observables are random variables sampled from that distribution.
41
2.1.2 Dynamics
Dynamics of a physical system describes how the system evolves in time. When we use
canonical variables tq, pu to describe the system, then dynamics is defined by the time
evolution of these variables
Bq BHpq, pq Bp BHpq, pq
“ “´ , (104)
Bt Bp Bt Bp
qpt0 q “ q0 ppt0 q “ p0 . (105)
where Hpq, pq is called the Hamiltonian of the system. If gradpHq is continuous, then the
solution to the Hamiltonian equations of motions exists; and if gradpHq is continuously
differentiable then the solution is unique. In this case, it can be extended to all times
t P R and the evolution tq, pu Ñ tqptq, pptqu is continuous in time t P R, and depends
continuously on the initial data tqpt0 q, ppt0 qu “ tq0 , p0 u.
The evolution of the canonical variables implies that the results of measurements
evolve in time. Since all observables are constructed from canonical variables, all ob-
servables evolve in time. In particular, this implies a one-parameter family of continuous
invertible mappings αt0 ,t : CpΓq Ñ CpΓq from the algebra of observables into itself so that
If f, g are two polynomials of canonical variables then it is easy to see that for β, γ P C
and every t P R we have,
By continuity, these properties extend to the whole algebra CpΓq, and thus αt0 ,t defines a
one-parameter family of ‹-automorphisms (‹-isomorphisms from the algebra into itself).
When the dynamics depends only on the difference t ´ t0 , i.e. it is time-translation invari-
ant, then the one-parameter family is actually a one-parameter group of automorphisms
tαt : t P Ru on CpΓq. By duality it defines a one-parameter group of transformations αt‹
on the state space as
(110)
` ˘
ωt pf q “ pαt‹ pωqqpf q “ ωpαt pf qq f P CpΓq .
42
In summary we see, that observables of a classical system define the commutative C ‹ -
algebra of observables CpΓq, and physical states define (algebraic) state on that algebra;
dynamics of the canonical system is given by a one-parameter group of ‹-automorphisms
on CpΓq. In particular, all physically observable features of a physical system are captured
by the algebraic formulation. However, still in the standard approach we start with the
phase space that describes our system, and then derive the C ‹ -formulation.
The Gelfand representation Theorem 1.46, however, allows us to reverse the logic.
Instead of starting with the phase space, and deriving the algebra from it, we can start
with a commutative C ‹ -algebra A that captures an abstract characterization of observ-
ables of our system; define the physical states as the state space SpAq; and define the
dynamics as a ‹-automorphism on A. By the Gelfand representation Theorem the ab-
stract algebra A can be represented as an algebra of continuous functions CpXq on a
compact Hausdorff space X. Every dispersion-free state of A is associated with the point
in X and so the space X gets the meaning of the phase space. In this sense, the algebra
comes first, from which the phase space is then derived. This is the starting point for
the algebraic construction. Using this motivation we aim in the next section to define a
general algebraic framework for physical theories.
States: The states of a physical system are described by rays (essentially unit vectors)
in a separable Hilbert space H.
Observables: The observables of a physical system are described by the set of bounded
self-adjoint operators in H.
43
A performed on the system in the state ω, is given by the Hilbert space matrix element
(111)
@ D
ωpAq “ ψ, Aψ
Dirac canonical quantization: Canonical coordinates and momenta satisfy the follow-
ing commutation relations
(112)
` ˘
rqi , qj s “ 0 “ rpi , pj s, rqi , pj s “ ı~δi,j , i, j “ 1, 2, . . . , n .
Bψ
(113)
` ˘
qi ψpxq “ xi ψpxq, pj ψpxq “ ´ı~ pxq i, j “ 1, 2, . . . , n .
Bxj
Even though the physical framework proved to be very successful, the physical basis
for many of its assumptions was unclear and unjustifiable. For that reason, a group
of mathematicians/physicists pioneered by Jordan, von Neumann, and Wigner, aimed
to provide a conceptually clear, and mathematically rigorous formalism for the above
postulates; to justify some of the tools in use of quantum mechanics; and to disregard
some of the assumptions not supported by mathematics or physical intuition. In this
section we discuss the summary of that approach that took a long time to be developed.
The idea of the algebraic approach is to use reasonable physical assumptions to define
an algebraic structure on the set of observables of a theory. Then to show that this
algebraic structure is rich enough to provide all the necessary tools used in quantum
mechanics as formulated in the above axioms. Following the literature, we formulate the
necessary assumptions on the set of observables as axioms.
@ D
Axiom 1. To each physical system Σ we can associate the triple pA, S, ; q formed by the set
@ D
A of all observables, the set S of all its states, and a mapping ; : pA, Sq Ñ R that assigns
@ D
to each pair pA, φq P pA, Sq a real number φ; A that we interpret as the expectation value
of the observable A in the state φ.
With a state of a physical system, we identify the method used for its preparation.
That is a (reproducible) way to prepare a configuration of the system. We assume that a
state is fully determined by all its expectation values, that is if φ, ϕ P S are such that
(114)
@ D @ D ` ˘
φ; A “ ϕ; A APA ,
then this must imply that φ “ ϕ. That is, we assume that there are no hidden properties
that we can never measure. We say the observables separate the states.
With an observable we identify a measurement device. Thus for any A P A we can
imagine a device with a pointer scale. The expectation value, is then the average over
44
n measurements of A in identically prepared states ω. Ideally, n shall go to infinity,
practically it has to be large enough to assure that our average is meaningful. If any
two devices produce the same expectation values for all states φ P S, then we identify
these devices. That is we assume that each measurement device produces a unique set
of expectation values, when evaluated in all states of S. We say, that S separates the
observables or simply that the set S is full with respect to A (see Figure 6 for intuition).
A S
A
B 1) switch on the stove
2) take 100g sugar
3) eat it
¨¨¨
C 4) hope for the best
..
. ..
.
Figure 6: The set of observables and the set of states of a physical system. After preparing
a state according to a given method we can measure it with all observables and identify it
with a list of all expectation values. If any to methods lead to the same list of expectation
values we identify these methods. On the other hand, if any two observables provide two
identical lists of expectation values when measuring all states, we identify them either.
Next we realize that we can use the states to compare observables. For A, B P A we
say that A ď B if φ; A ď φ; B for all state φ P S. In particular, we call A ě 0 positive
@ D @ D
if φ; A ě 0 for every state φ P S. Especially, this implies that the states are positive
@ D
functionals on A.
Axiom 2. The relation ď defined above is a partial ordering on A; in particular A ď B and
B ď A imply A “ B.
With our interpretation of observables as measurement devices we see that the con-
dition φ; A ě 0 for every φ P S implies that every measurement A is positive. Since
@ D
the set of state is full with respect to the set of observables this meas that the scale of
the measurement device represented by A is positive. Thus we can identify the positive
element with measurement devices whose pointer scale ranges in positive real numbers.
Next we notice that for any two observables A, B P A the set t φ; A ` φ; B : φ P Su
@ D @ D
45
Axiom 3.
i ) For each observable A P A and any λ P R there exists an element pλAq P A such that
(115)
@ D @ D ` ˘
φ; λA “ λ φ; A φPS .
(116)
@ D @ D @ D ` ˘
φ; A ` B “ φ; A ` φ; B φPS .
Axiom 4. For every observable A the set S contains the set of dispersion-free states SA for
A that uniquely determines the one-dimensional subspace of A generated by A; for any two
observables A, B the dispersion-free states satisfy SA X SB Ď SA`B and SI “ S.
In particular, this axiom implies that for two observables A and B if SA Ď SB and
φ; A “ φ; B for every φ P SA then A is equal to B. Thus dispersion-free states in a
@ D @ D
certain sense, uniquely determine the observables. This reflects our original motivation
to introduce this axiom. Nevertheless, this axiom requires an assumption of dispersion-
free states, which is an idealization of a physical process of measurement, as we dis-
cussed above. To argue that such states are reasonable, we can take the standpoint that
even though every realistic measurement is noisy, the noise can be arbitrary reduced.
Therefore there is no fundamental (nature given) threshold which we cannot surpass
by making a measurement more precise/less noisy. (In fact, this may be violated if one
takes General Relativity into account. But we are not going into this discussion here.)
46
If A is an observable that represents a measurement device then we can interpret A2
as a measurement device whose scale is the square of the pointer scale of A. In particular,
A2 is therefore always positive. Similarly, we can define any power An for a non-negative
integer n, and set I “ A0 . We put the existence of such powers as an additional axiom.
Axiom 5. For any observable A and any non-negative integer n there exists at least one
element An P A such that
i ) SA Ď SAn ,
@ D @ Dn
ii ) φ; An “ φ; A for all φ P SA .
From the second point of this Axiom and Axiom 4 we see, that the element An is
unique in A and we can unambiguously call it the nth power of A.
With this axiom the vector space A obtains a power structure. Moreover, it follows
that A contains a unit I “ A0 for every A P A and an element O such that
and (117)
@ D @ D ` ˘
φ; I “ 1 φ; O “ 0 φPS .
From the second point of this axiom, it follows that SI “ SA0 “ S, and that the usual
rules apply: A0 “ I, A1 “ A, I n “ I and On “ O for n ą 1.
Furthermore, since φ; A2 ě 0 for all φ P SA and since SA fully determines A2 it
@ D
Stated more precisely, we never defined the product between A and B that again defines
an observable AB. Notice, that in the usual approach to quantum mechanics, where we
identify observables with self-adjoint operators on a Hilbert space this is in particular
false: a product of self-adjoint elements is in general not self-adjoint.
On the other hand, for any observable A the product AA is defined by the unique
element A2 from the Axiom 5. We use this observation to equip A with a product.
47
This product depends on powers of the same observable (which we defined) and on
sums between observables (which we also defined). Moreover, this product is commuta-
tive, A ˝ B “ B ˝ A, and from Axiom 4 and 5 it satisfies
(120)
@ D @ D@ D ` ˘
φ : A ˝ B “ φ; A φ; B φ P SA X SB .
Thus, since dispersion-free states uniquely determine the observables it follows that An ˝
Am “ An`m . In particular, we have A ˝ I “ A. In other words, the power structure that
we had derived after the previous Axiom, can be generated by the symmetric product
on A. Therefore ˝ naturally extends the structure that we already have. This makes A
almost an algebra with identity, however, not quite yet.
The symmetric product, even though commutative, is neither associative, pA˝Bq˝C ‰
A˝pB ˝Cq, nor does it distribute over addition. The non-associativity is not a big loss, and
in general there is little algebraic reason to want to restore associativity. Distributivity, on
the other hand, is crucial for a definition of an algebra. At the same time, it seems hard
to find a physically compelling argument to make the product distributive. However,
one can show (even though we do not do it here) that distributivity follows from the
following mild looking assumption.
Axiom 6. For any pair of observables A, B the symmetric product is homogeneous in the
fist entry, i.e. for every real number λ it satisfies λpA ˝ Bq “ pλAq ˝ B.
Since the product is commutative, homogeneity in the first entry implies homogeneity
in both entries. With this assumption the symmetric product becomes distributive. In
fact, it also becomes weakly associative meaning that for A, B P A the product satisfies
A ˝ pB ˝ A2 q “ pA ˝ Bq ˝ A2 . A commutative algebra with a symmetric product that
satisfies the weak associativity is called a Jordan algebra; and a Jordan algebra that
satisfies the reality condition is called a real Jordan algebra. Therefore we can say that
with the first 6 Axioms the set of observables A becomes a real Jordan algebra. Notice,
that the Axioms 1-5 are very natural from the physicists perspective. The Axiom 6 is
the only questionable one. However, even though there is no direct evidence that the
product must be distributive, there is no direct physical reason why it should not be.
Mathematically, this assumption makes a significant difference.
At this point, our set A is an algebra, but it remains to be shown that this algebra
is indeed useful: that it provides a computational or conceptual advantage over the
standard approach to classical or quantum theory. In particular, it is important to classify
such algebras, and to understand their representations. To approach this question we
need to introduce topology on A.
(121)
@ D
}A} “ supt| φ; A | : φ P Su.
48
Identifying observables with measurement devices, relation (121) assigns the largest
possible expectation value to each measurement. This mapping has almost all the prop-
erties of a norm: if }A} “ 0 then φ; A “ 0 for every φ P S, and since S separates the
@ D
observables we get A “ 0; by Axiom 3 each state is linear and thus, for A P A and λ P R
we have }λA} “ |λ|}A}; with B P A it is clear that }A ` B} ď }A} ` }B}. Moreover,
from the reality condition it also holds that }A2 } ď }A2 ` B 2 } for any A, B P A. The only
missing requirement such that (121) defines a norm, is that it is finite for every element
of A. We put this as an axiom.
Axiom 7. Each element A in A has finite norm and A is complete in the norm topology,
@ D
and we identify elements in S with positive continuous linear functional with φ; I “ 1.
The reason why we require the norm of observables to be bounded can be again un-
derstood with our identification of observables with measurement devices (even though,
by now the space A is significantly larger than it was in the beginning of our discussion).
Each practical measurement device has a finite (bounded) scale. Therefore, if A repre-
sents a measurement device then even the largest expectation value of A is going to be
bounded by the scale of that device. Measurements that contain infinity as a valid mea-
surement outcome are even theoretically not feasible. Therefore this axiom is reasonable
on physical grounds. Assuming that A is complete, on the other hand, is merely a practi-
cal requirement. Even though it does enlarge the space A, it does not significantly alter
the physical interpretation.
With this norm each state in S is by definition continuous. Thus a state φ P S is a
positive, linear, continuous functional on the algebra A with φpIq “ 1. Thus, S is the space
of states in the sense of our algebraic Definition 1.42 ii ). For that reason, henceforth we
will write SpAq instead of S and use our standard notation ωpAq instead of ω; A .
@ D
Proof. By definition of the norm it follow that ωpAq ď }A} for every state ω P SpAq.
Thus by linearity of ω we have ωp}A}I ˘ Aq ě 0, so both }A}I ˘ A are positive. Thus
p}A}I ´ Aqp}A}I ` Aq is also positive (e.g. because it is uniquely determined by the
dispersion-free states/ two measurement devices both with positive pointer scale). Then
for every state ω P SpAq,
(123)
` ˘
}A}2 ´ ωpA2 q “ ω p}A}I ´ Aqp}A}I ` Aq ě 0,
49
On the other hand, the positivity of p}A}I ˘ Aq2 “ }A}2 ` A2 ˘ 2}A}A implies that
for any state ω we have
From the Proposition 1.50 we know that states induce an inner product on the vector
space on which they are defined; in this case on A by,
(125)
@ D ` ˘
ωpA ˝ Bq “ A, B A, B P A .
Moreover, we have seen that this product satisfies the Cauchy-Schwarz inequality. This
inequality, in turn, implies that the product is continuous since for A, B P A we have
for every ω P SpAq and every pair A, B P A. Taking the supremum on both sides of (126)
and using Lemma 2.3 implies }A ˝ B}2 ď }A2 }}B 2 } “ }A}2 }B}2 for every A, B P A.
A real Jordan algebra, which is a Banach space in the norm in which the product is
continuous is called a Jordan-Banach algebra. However, we have more structure. The
norm of our Jordan-Banach algebra also satisfies }A2 } “ }A}2 and the reality condition
}A2 } ď }A2 ` B 2 }. Such Jordan-Banach algebras are called JB-algebras.
In summary, A is a JB-algebra which is commutative, distributive, but not-necessarily
associative with a continuous product. From the mathematical point of view, however,
it is easier to treat an algebra which is associative but not-necessarily commutative. Be-
cause of this, it is convenient to ask the question when the JB-algebra can be embedded
into a larger algebra such that the symmetric product is given by the (associated but not
necessarily commutative) product AB such that
1` ˘ AB ` BA
A˝B “ pA ` Bq2 ´ A2 ´ B 2 “ . (127)
2 2
To achieve this, we can first complexify the algebra A by considering the set
tA ` ıB : A, B P Au, (128)
and then introduce the involution by ‹ : A ` ıB Ñ A ´ ıB. One can show that most of the
properties of the C ‹ -algebra are satisfied. Especially, one can extend the above norm such
that it satisfies the C ‹ -property }A‹ A} “ }A}2 and that }A‹ } “ }A}. However, the product
on this algebra may still be non-associative. This is the reason why the embedding may
sometimes fail.
A JB-algebra is called special if its embedding into a C ‹ -algebra exists; otherwise it
is called exceptional. Even though there is always a homomorphism from a JB-algebra
into a C ‹ -algebra [11, thm. 7.1.8], there may not exist an isometric isomorphism that
50
embeds a JB algebra into a C ‹ -algebra. A JB-algebra that is isometrically isomorphic to
a Jordan subalgebra of a C ‹ -algebra is called a JC-algebra. The following theorem [11,
thm. 7.2.3 and lem. 7.2.2] characterizes all exceptional JB algebras.
Theorem 2.4. Any exceptional JB-algebra A contains a unique purely exceptional ideal I
such that A{I is a JC-algebra. The ideal I is itself a Jordan algebra and each of its factor
representations is onto H3 pOq.
In the above Theorem H3 pOq denotes the algebra of 3 ˆ 3 self-adjoint matrices over
octonions. This algebra seems too poor for physical applications, and so an embedding
of a JB-algebra into a C ‹ -algebra is quite natural. We state this as a finial axiom.
Axiom 8. The algebra of observables A can be identified with the set of all self-adjoint
elements of a not necessarily commutative C ‹ -algebra A.
To conclude: We started with a set of observables A and states S; under mild physi-
cal assumptions we derived an algebraic structure on A. We further realized, that the
resulting algebra can be identified with a set of self-adjoint elements of a C ‹ -algebra A.
Now we have the entire algebraic machinery for C ‹ -algebras. In particular, by means
of the GNS construction we know that each state ω on A yields a cyclic representation
pπω , Hω , xω q. In particular, by the ‹-homomorphism property self-adjoint elements of A
are mapped to self-adjoint operators on the Hilbert space Hω . By the “Expectation” axiom
of Dirac-von Neumann and the unitarily equivalence clause of the GNS Theorem, we see
that the framework of quantum mechanics is unitarily equivalent to the so derived GNS
representation. Thus, we recovered the first three Dirac-von Neumann axioms.
The first three axioms of Dirac-von Neumann are of general nature. They do not refer
to a quantum system, but describe a physical system in general. The fourth and the fifth
axioms, on the other hand, refer specifically to a quantum system. It is for this reason
that with our general construction we recovered the general axioms. As we will see in
the next section, the fourth Dirac-von Neumann axiom will guide us to define an explicit
C ‹ -algebra for quantum mechanics. The fifth axiom drops entirely by the result of the
Stone-von Neumann theorem.
51
Additionally we notice that in the algebraic formalism, the structure of quantum me-
chanics becomes derived, just like the structure of the phase space of classical mechanics
was derived from the commutative C ‹ -algebra. In particular, we see that the Hilbert
space of quantum mechanics is not a fixed construction of the framework, but a property
that arises from a choice of a physical state, i.e. from the method of preparation of a
physical system. Different methods of preparation may lead to different Hilbert spaces.
Indeed, in the standard approach to quantum mechanics, such transitions between states
that lead to different Hilbert spaces typically result in exploding physical quantities such
as the magnetization, susceptibility, conductivity, and the like. If this happens, we say
that a system undergoes a phase transition. Such an explosion of physical quantities
results from the fact, that the scalar product on the originally defined Hilbert space be-
comes inappropriate for the new physical system, since for the new system a new Hilbert
space is required. In the algebraic approach, such s transition corresponds to the change
between states ω1 and ω2 that lead to two not unitarily equivalent GNS representations
of the C ‹ -algebra. In this way, one avoids divergent quantities and can describe phase
transitions in a mathematically careful way.
52
3 A quantum mechanical particle
Section remarks: This section is based on several books and papers. The semi-
historical note is motivated by the discussion in the Heisenberg’s book [12] and his
papers [13, 14]. The uniqueness of the Weyl algebra is from [10]. The proof of the
Stone-von Neumann theorem and the Schrödinger representation of the Weyl algebra
is adopted from [6] and the original paper of von Neumann [15].
After the last section we know that we need a C ‹ -algebra of observables to define a
physical system. For classical physics the algebra has to be commutative. In this case,
we can choose the algebra of continuous functions on the phase space. The question is,
what C ‹ -algebra do we choose to describe quantum mechanics. To answer this question
we use the canonical commutation relations and the Heisenberg uncertainty principle.
Before we proceed, we mention a semi-historical note that shall motivate our approach.
53
be
~
∆x∆p ě , (130)
2π
where ∆x and ∆p denote the variance of position and momentum measurements. He
suggested that this uncertainty is a fundamental principle of nature; it is impossible
to measure position and momentum of an electron with better simultaneous precision.
Today we call it the Heisenberg’s uncertainty principle.
To describe this phenomenon quantitatively, he suggested to use the wave picture
of the electron: the electron is a wave packet, electron’s position is the localization of
the packet, and electrons velocity is the group velocity of the packet. The sharper the
wave packet is localized in space the more spread it is in momentum, thereby causing
the packet to disperse in time. Thus, a sharply peaked wave packet quickly spreads
and its group velocity is not clearly defined. The analysis of the problem requires care,
because the aforementioned wave packet describes the probability distribution for the
electron and not the electron itself. Moreover, the dispersion of the packet depends on
the implemented dynamics, as for non-linear processes the wave package may keep its
shape for an extended time period. Nevertheless, the above heuristic argument already
captures the main idea of the rigorous calculation.
Heisenberg’s idea was groundbreaking and is considered as the initial formulation of
quantum mechanics. Nevertheless, it was Born and Jordan who realized that Heisen-
berg defined observables as infinitely large matrices, i.e. linear operators on an infinitely
dimensional Hilbert space [16]. In particular, being operators, observables may not al-
ways commute. Using Heisenberg’s uncertainty principle Born and Jordan derived that
the position and the momentum of an quantum particle must satisfy what we nowadays
know as the canonical commutation relations or CCRs,
(131)
` ˘
rxi , xj s “ 0 “ rpi , pj s, rxi , pj s “ xi pj ´ pj xi “ ı~1, i, j “ 1, . . . , d ,
where d is the number of degrees of freedom for the electron (e.g. a free particle moving
in a d-dimensional space; no spin is considered); and xj (resp. pj ) denotes the position
(resp. momentum) observable in the jth direction. Often we define x “ px1 , . . . , xd q and
p “ pp1 , . . . , pd q as operators on d copies of the Hilbert space H, and phrase (131) as a
single CCR relation
In summery, the Heisenberg’s uncertainty principle implies that the position and the
momentum observables of an electron do not commute. Moreover, the CCRs (131) imply
that either of the operators, xi or pi (or both), must be unbounded for every i “ 1, . . . , d.
54
To see this, we write the CCRs (dropping the index for clarity) as xp “ ı~ ` px. Then,
The above relation implies that the norms of the operators x and p satisfy
Now, }xn´1 } cannot be zero, otherwise 0 “ }xn´1 } “ }x}n´1 would imply x “ 0 which
violates the CCR. Thus, (134) reduces to
n
(135)
` ˘
~ ď }q}}p} nPN .
2
Because n can be arbitrary large, at least one of the operators has to have an unbounded
norm. This, in turn, implies that we have to be careful with the multiplication of opera-
tors xi and pi . In particular, this multiplication can be defined only on an empty set —
implying that it is not meaningful. Moreover, even if the multiplication can be defined on
the dense subset of the Hilbert space, it cannot be defined on the whole space. Because of
this, the whole concept of “commutation” for unbounded operators is non-trivial. Many
technical complications arise when we want to formulate CCRs using clean mathematics.
We now return to our algebraic approach. In the algebraic approach, the mathemat-
ics is simple because all elements have finite norm. We only need to define what algebra
is suitable to define observables of a particle. Ideally, we would like to use x and p to
motivate the algebra, but since these operators are unbounded, they cannot generate a
C ‹ -algebra. Despite this, we can use x and p in a heuristic way (not rigorous calcula-
tions), to get an intuition for the right C ‹ -algebra. One way to do this was suggested by
Weyl and it leads us to the definition of the Weyl algebra.
55
Since the concept of “commutation” is complicated for unbounded operators, Weyl
suggested to look at the operators
(136)
` ˘
U pαq “ eıαx V pβq “ eıβp α, β P Rd ,
where we use the abbreviation αx “ di“1 αi xi and βp “ di“1 βi pi . The operators U pαq
ř ř
and V pβq are called the Weyl operators, and the vectors α and β label these operators.
Since x and p suppose to represent observables we assume them to be self-adjoint. In
this case, the Weyl operators are bounded and can be defined via the spectral calculus.
We want to use the Weyl operators to generate our algebra of observables. For this we
need to define a product between them. To this end we use the CCRs and the Baker-
Campbell-Hausdorff formula.
Proof. For the proof and a more general statement consult [17, p. 25 prop. 2].
Formally, the CCRs say that rx, ps is proportional to the identity; thus it commutes with
every operator, and we can use the Baker-Campbell-Hausdorff formula to define the
following replacement of the CCR for the Weyl operators,
U pαqV pβq “ V pβqU pαqe´ı~αβ , U pαqU pβq “ U pα ` βq, V pαqV pβq “ V pα ` βq, (138)
where αβ “ di“1 αi βi is the Euclidean scalar product. These relations are called the
ř
commutation relations in the Weyl form or simply the Weyl relations. Even though U pαq
and V pβq together with (138) is enough to proceed, it will be more convenient in the
following to work with a single operator instead of two. To this end, we use vectors
in Rd ˆ Rd , such that the first d-components represent the vector α, and the second d-
components represent the vector β. Then for v “ pα, βq and w “ pγ, δq in Rd ˆ Rd p“ R2d q
we define the following operation,
˜ ¸˜ ¸˜ ¸
α 0 1 γ
σpv, wq “ αδ ´ γβ “ . (139)
β ´1 0 δ
Especially, σpv, vq “ 0 for every v P R2d . The mapping σ : R2d ˆ R2d Ñ R is a particular
case of what is called a non-degenerate symplectic bilinear form. It has a profound meaning
in classical mechanics. Formulating our C ‹ -algebra using σ is thus preferable as it can
give us a general procedure how to define a quantum algebra from the classical case. In
other words: how to quantize a classical system.
56
Indeed, it turns out that if we use v “ pα, βq P R2d and set
αβ αβ
W pvq “ e´ı~ 2 V pβqU pαq “ eı~ 2 U pαqV pβq. (140)
then the Weyl relations can be derived from the following relations
σpv,wq
W pvq‹ “ W p´vq, W pvqW pwq “ W pv ` wqe´ı~ 2 . (141)
In the following, we will not use U pαq and V pβq but instead only the W pvq’s labeled by
vectors in R2d . For this reason we will call W pvq the Weyl elements and refer to (141)
as to the Weyl relations. Notice, however, that we can recover the operators U pαq and
V pβq by using vectors of the form v “ pα, 0q P R2d or v “ p0, βq P R2 , respectively. Even
though, the above calculations were formal, we can now use (141) as a starting point to
define a clean C ‹ -algebra.
We start with the set of abstract elements tW pvq : v P R2d u, that transform under a ‹-
operation and satisfy the multiplication relations according to (141). Then we consider
an abstract algebra AW generated by this set. That is AW comprises elements of the form
for a1 , . . . , an P C and v1 , . . . , vn P R2d . Notice that due to the multiplication relations, all
monomials of Weyl operators can be rewritten as a single Weyl operator multiplied by a
complex number. To form a C ‹ -algebra, we need to equip AW with a norm that satisfies
the C ‹ -property. But what norm shall we choose? The following theorem provides an
answer to this question. For the proof see [10, thm. 5.2.8].
Theorem 3.1. Let V be a real linear space equipped with a non-degenerate symplectic
bilinear form σ. Let A1 and A2 be two C ‹ -algebras generated by nonzero elements Wi pvq,
v P V , satisfying (141). Then there exists a unique ‹-isomorphism α : A1 Ñ A2 such that
(143)
` ˘
α W1 pvq “ W2 pvq, pv P V q.
Proposition 3.2. Let W be the Weyl algebra generated by elements tW pvq : v P R2d u that
satisfies the Weyl relations. Then
i ) W is unital and W p0q “ I is the unit of W;
ii ) for each v P R2d the element W pvq is unitary;
iii ) for each non-zero v P R2 the spectrum of W pvq is the entire unit circle of C.
57
Proof. For i ), observe from the Weyl relations that
(144)
` ˘
W pvqW p0q “ W pvq “ W p0qW pvq v P R2d .
Since each element in W is a limit of finite complex linear combinations of the Weyl
elements, for every A P W we have AW p0q “ W p0qA “ A, and thus W p0q “ I is the unit
of the Weyl algebra; in particular, W is unital.
For ii ), we use again the Weyl relations to see that
v P R2d . (145)
` ˘
W pvqW pvq‹ “ W pvqW p´vq “ I “ W p´vqW pvq “ W pvq‹ W pvq
The right hand side does not have a two-sided inverse, and consequently eıα λ is in the
spectrum of W pvq. It follows that the spectrum is invariant under rotation and since λ is
an element of the unit circle, the spectrum must be the entire unit circle in C.
Remark. The Theorem 3.1 holds even if V is infinite dimensional. In this case,
the Weyl algebra describes a system with infinitely many degrees of freedom. Such
a theory is called (non-relativistic) quantum field theory. Opposed to it, quantum
mechanics describes systems with a finite number of degrees of freedom and thus
the Weyl algebra is labeled by a finite dimensional V .
To conclude: Using the canonical commutation relations that follow from the Heisen-
berg uncertainty principle we motivated a C ‹ -algebra for the quantum particle. This
algebra is essentially unique and we call it the Weyl algebra W. To connect our algebraic
formalism to the usual (Schrödinger) quantum mechanics, we need to represent the al-
gebra of observables as bounded linear operators on a Hilbert space. As we have seen, we
can use the GNS construction for this. Nevertheless, we have also seen, that not all rep-
resentations are necessarily unitarily equivalent. And if the Weyl algebra has different
unitarily inequivalent representations, then these representations will yield physically
different descriptions of the quantum particle. Which one should we then choose? In
particular, if different inequivalent representations exist, what is their physical signifi-
cance? And then, how many different inequivalent representations of the Weyl algebra
can we construct. All these questions were answered by von Neumann. The answer is
known as the Stone-von Neumann theorem.
58
3.3 Stone-von Neumann theorem
We now want to investigate unitarily inequivalent representations of the Weyl algebra.
Recall that, unitarily inequivalent representations lead to different Hilbert spaces on
which we can represent the C ‹ -algebra as a (sub)algebra of bounded operators. If the C ‹ -
algebra suppose to define a physical system, then the Hilbert spaces (especially their inner
products) define the expectation values. Thus they are physically relevant. If several
inequivalent representations exist then we need to select the one which is best suited for
the description of the system.
Recall, that irreducible representations are the building blocks of all representations.
Thus, we only need to characterize irreducible representations of the Weyl algebra. We
can restrict the class of “interesting” representations even further by requiring a mild but
physically well motivated regularity condition. To formulate this condition we need to
refresh the definition of the strong operator topology.
Definition 3.3 (Strong operator topology). Let BpHq be the set of bounded linear
operators on the Hilbert space H. For every x P H define a map from BpHq into H by
A Ñ Ax. (147)
The strong operator topology is the weakest topology such that each of these maps is
continuous. In this topology, a sequence of operators pAn q converges to an operator A,
if the sequence of vectors pAn xq converges to the vector Ax for every x P H. It is a
consequence of the uniform boundedness principle that a strong limit of a sequence of
bounded linear operators is again a bounded operator.
Definition 3.4. An operator-valued function U ptq defined on R and acting on the Hilbert
space H is called a strongly continuous one-parameter unitary group if
i ) for each t, s P R, the operator U ptq is unitary, and U pt ` sq “ U ptqU psq;
ii ) and for x P H, the vector U ptqx converges to the vector U pt0 qx as t Ñ t0 .
We mention some well known results of functional analysis to point out the impor-
tance of the strong operator topology.
Stone’s theorem. We state the theorems and refer to e.g. [3, thm. VIII.7 and thm.
VIII.8] for a more complete statement and the proofs.
(148)
` ˘
U ptq “ eıtA tPR ,
59
` ˘ ` ˘
the limit lim 1t U ptqx ´ x exists and is equal to ıAx; conversely, if lim 1t U ptqx ´ x
tÑ0 tÑ0
exists, then x P dompAq.
For v P R2d of the form vn “ p0, . . . , 0, t, 0, . . . , 0q P R2d (where all except the nth ele-
ment are zero), the Weyl operators define one-parameter unitary groups. In conjunction
with the Stone’s theorem, the regularity condition assures that the operators π W pvn q
` ˘
will have generators. Hence, we have a chance to define positions and momenta from
the Weyl algebra. Since this is desirable on physical grounds, we restrict our attention
only to regular representations. Thus we aim to characterize all regular irreducible rep-
resentations of the Weyl algebra.
Remark: The regularity condition says that if pvn q is a sequence in R2d that con-
verges to a vector v, then the sequence of operators pπ W pvn q q will strongly con-
` ˘
verge to an operator π W pvq . Notice, that this is not necessarily the case in the
` ˘
C ‹ -norm since
` σpvn ,vq ˘ σpvn ,vq
}W pvn q ´ W pvq}2 “ }W p´vq eı~ 2 W pvn ´ vq ´ I }2 “ }eı~ 2 W pvn ´ vq ´ I}2 .
Assume that vn ´ v ‰ 0 for all n P N, then by Proposition 3.2 iii ), the spectrum of
W pvn ´ vq is the entire unit circle in C and we have
σpvn ,vq
}eı~ 2 W pvn ´ vq ´ I} ě supt|λ ´ 1| : λ P C, |λ| “ 1u “ 2. (149)
60
with respect to the measure µ on R2d , then the operator
ż
A“ f pvqU pvq dµpvq, (150)
R2d
can be defined as a limit of Riemann sums. We will use this construction in the following
proof, therefore we review it in some detail now.
Pn
Pm
C2
C1
Integrals via Riemann sums. For a bounded set O Ă R2d we write Pn pOq “
tC1 , . . . Cn u to denote a partition of O into n (open) cells Ci Ă O. For two partitions
P1 pOq and P2 pOq we write P2 pOq ě P1 pOq and say that “P2 pOq is finer than P1 pOq” if
each cell C of P1 pOq partitions into a collection P2 pCq of cells in P2 pOq such that (see
Figure 7)
ď ÿ
C“ K µpCq “ µpKq. (151)
KPP2 pCq KPP2 pCq
Given a partition Pn pOq, choose one vector vC from each cell C and define an operator
ÿ
An “ f pvC qU pvC q µpCq. (152)
CPPn pOq
61
Then for any P1 pOq ě P0 pOq we have
› ÿ ÿ ›
}A0 x ´ A1 x} “ › f pvC qU pvC qµpCq ´ f pvK qU pvK qµpKq›
› ›
CPP0 pOq KPP1 pOq
› ÿ ÿ ` ˘ ›
f pvC qU pvC q ´ f pvK qU pvK q µpKq›
› ›
“›
CPP0 pOq KPP1 pCq
ÿ ÿ › ›
f f K ›µpKq
› ›
ď › Cpv qU pvC q ´ pvK qU pv q
CPP0 pOq KPP1 pCq
ď ,
where the second equality holds by (151) since P1 pOq is finer than P0 pOq. Thus, if
Pn pOq is an increasing sequence of partitions, i.e. Pn pOq ą Pm pOq when n ą m,
` ˘
then for every x P H there is a k P N for which Pk pOq is fine enough to satisfy (153)
and the sequence of vectors pAn xq is Cauchy; its limit defines a linear mapping AO
such that,
ż
(154)
` ˘
x ÞÑ lim An x “ f pvqU pvqx dµpvq “: AO x xPH .
nÑ8 O
Since the strong limit of bounded linear operators is bounded and thus AO defines
the integral of f U over O.
To see that the limit does not depend on the partition sequence, choose two
increasing sequences Pn pOq and P̃n pOq . Then there is a sequence P n pOq that is
` ˘ ` ˘ ` ˘
finer than both Pn pOq and P̃n pOq; for example define the cells of P n pOq as non-empty
intersections of each cell of Pn pOq with each cell of P̃n pOq. These three partition
sequences give rise to three sequences of operators pAn q for the partitions Pn pOq ,
` ˘
pBn q for the partitions P̃n pOq , and pCn q for the partitions P n pOq . From the above,
` ˘ ` ˘
Because f is integrable, the right hand side of the above estimate goes to zero as m
goes to infinity. Therefore the sequence of integrals over increasing balls is a Cauchy
62
sequence of operators. The limit point of that sequence defines our desired integral
ż
A“ f pvqU pvq dµpvq. (155)
R2d
Notice that he matrix elements of A are the usual (Riemann) integrals of matrix
elements of U pvq since matrix elements of Riemann sums are the Riemann sums of
matrix elements. Explicitly, for x, y P H we have
ż
(156)
@ D @ D
x, Ay “ f pvq x, U pvqy dµpvq.
R2d
Indeed we could use (156) as a definition for the operator A in terms of its matrix
elements. This construction of integral operators is often useful when we need to
define operators that are invariant under certain transformations. For this one has
to find the integral kernel f pvq such that the effect of the transformation can be
“averaged out”.
Now, we go back to the regular irreducible representations of the Weyl algebra. The
question is how many regular irreducible unitary inequivalent representations does a
Weyl algebra have. The Stone-von Neumann theorem provides the answer. To simplify
the notation, for the rest of this section we will assume ~ “ 1, which is always possible
to do by suitably redefining the units of meter, second, and kilogram. This is why ~ will
not appear in our commutation relations in the following.
Theorem 3.8 (Stone-von Neumann). All regular irreducible representations of the Weyl
algebra are unitarily equivalent.
Before starting the proof notice that each state on the Weyl algebra is uniquely defined
by its action on the Weyl elements. To see this, observe that each element A of the Weyl
algebra W is by definition a norm limit of linear combination of Weyl elements of the
form (142) such that
ÿ
A“ an W pvn q. (157)
nPN
and ωpAq is uniquely defined by the expectation values on the Weyl elements.
Proof of Theorem 3.8. The idea of proof is to show that if π is a regular irreducible rep-
resentation of W on the Hilbert space H then there exists a vector xF P H such that
2
(159)
@ ` ˘ D
xF , π W pvq xF “ e´|v| {4 .
63
H1
BpH1 q xF |v|2
xF , π1 W pvq xF “ e´ 4
@ ` ˘ D
π1
H2
|v|2
yF , π2 W pvq yF “ e´ 4
@ ` ˘ D
π2 ` ˘ |v|2
W BpH2 q yF ωF W pvq “ e´ 4
H3
|v|2
π3 @
zF , π3 W pvq zF “ e´ 4
` ˘ D
BpH3 q zF
Then such a vector defines an (algebraic) state that is uniquely specified by the relation
2
(160)
` ˘ @ ` ˘ D ` ˘
ωF W pvq “ xF , π W pvq xF “ e´|v| {4 v P R2d ,
all v P R2d such that supt}π W pvq } : v P R2d u ď 1. Since the function v Ñ e´}v} {4 is
` ˘ 2
continuous and integrable on R2d the conditions of our integral construction are satisfies
and we can define the operator
ż
1 |v|2
(161)
` ˘
P“ e´ 4 π W pvq dµpvq.
2π R2d
64
For x, y P H we have
ż
1 ´|v|2
(162)
@ D @ ` ˘ D
x, Py “ e 4 x, π W pvq y dµ
2π R2d
ż
1 ´|v|2
(163)
@ ` ˘ D
“ e 4 π W p´vq x, y dµ
2π R2d
ż
1 ´|v|2
(164)
@ ` ˘ D @ D
“ e 4 π W pvq x, y dµ “ Px, y ,
2π R2d
implying that P is self-adjoint. Moreover, P is not identically zero, otherwise for every
w P R2d and x, y P H we had
ż
1 |v|2 @
(165)
@ ` ˘ ` ˘ D ` ˘ D
0 “ π W pwq x, Pπ W pwq y “ e´ 4 x, π W pvq y eıσpv,wq dµpvq.
2π R2d
Writing v “ pα, βq and using the definition of σ we see that the right hand side of the
|v|2 @
above equation is the Fourier transform of the element e´ 4 x, π W pvq y . Thus (165)
` ˘ D
would imply that for every x, y P H the matrix element x, π W pvq y must vanish for
@ ` ˘ D
Next, we notice that by looking at the matrix elements Px, π W pzq Py for every
@ ` ˘ D
Setting z “ 0 we get
P2 “ P. (167)
To finish the proof, we need to show that such an xF is unique, that is that the range
of P is one-dimensional. Assume this is not the case, then there is a vector yF P H in the
range of P (i.e. PyF “ yF ), and orthogonal to xF . Then,
@ ` ˘ D @ ` ˘ D ´|v|2 @ D
yF , π W pvq xF “ yF , P π W pvq PxF “ e 4 yF , xF “ 0,
65
for every v P R2d , and thus by continuity yF , πpAqxF for every A P W. If yF ‰ 0, then
@ D
yF is in the orthogonal complement of the closure of the set of vectors tπpAqxF : A P Wu.
In particular, this set is a proper subset of H, invariant under the action of πpWq. This,
however, contradicts irreducibility of π. Therefore, yF “ 0.
The state ωF from the proof of the Stone-von Neumann Theorem is called the Fock
state. It is almost exclusively the only state that we use for explicit calculations in physics.
The GNS representation induced by the Fock state is called the Fock representation
and the GNS Hilbert space is (called) the Fock space.
66
formulation from fundamental physical principles and rigorous mathematical analysis.
In particular, we know that the Schrödinger quantum mechanics is a representation of
an underlying algebra that describes a quantum particle. In this section, we only show
the general concepts, as we do not have time develop the details of the construction.
We begin with the Hilbert space L2 pRd q of square integrable functions with respect to
the Lebesgue measure. In the following, it will be convenient to get back to the original
Weyl operators U and V . So with v of the form v “ pα, 0q P R2d we define U pαq “ W pvq
and with w of the form w “ p0, βq P R2d we define V pβq “ W pwq. Now we define a
representation of the Weyl algebra. That is, we define the action of the Weyl operators
on the Hilbert space L2 pRd q. For ψ P L2 pRd q we set
(170)
` ˘ ` ˘
U pαqψ pxq “ eıαx ψpxq, V pβqψ pxq “ ψpx ` βq,
where the equality is meant to hold almost everywhere. After analyzing this action we
arrive at the conclusion that U pαq and V pβq define a strongly continuous irreducible
representation of the Weyl algebra.
Using the Stone theorem, we can derive the operators U pαq and V pβq with respect to
their parameters. The derivatives at the origin yield the generators of the Weyl operators
in the Schrödinger representation. We can explicitly calculate,
´ BU pαq ˇˇ ¯ ´ BV pβq ˇˇ ¯ ´ B ¯
´ı ˇ ψ pxq “ xψpxq ´ı ˇ ψ pxq “ ´ ı ψpxq, (171)
Bαi α“0 Bβi β“0 Bxi
for i P t1, . . . , du. For each i we call the generator of U pαi q the position operator qi and
group all these generators as q “ pq1 , . . . , qd q so that U pαq “ eıαq . Similarly, for each i we
call the generator of V pβi q the momentum operator pi and group them to p “ pp1 , . . . , pd q
so that V pβq “ eıβp . If we use our notation id pxq “ x to denote the identity function and
∇ to denote the Euclidean gradient on Rd then (171) defines the action of q and p on
L2 -functions as
qψ “ id ¨ψ pψ “ ´ı∇ψ. (172)
These operators map L2 -functions into d-copies of L2 , that is into the space of square
integrable vector fields L2 pRd ; Rd q. In particular, we see that neither q nor p are bounded.
Their domains of definition follow from the Stone’s theorem such that
dompqq “ tψ P L2 pRd q : xψ P L2 pRd ; Rd qu, (173)
domppq “ tψ P L2 pRd q : ∇ψ P L2 pRd ; Rd qu “ H 1 pRd q, (174)
where H 1 pRd q denotes the Sobolev space of square integrable functions, whose deriva-
tives are also square integrable. Especially, on these domains q and p are self-adjoint.
Their product can be defined on the common core domain of rapidly decreasing func-
tions — the Schwarz space SpRd q — and on that space we can easily calculate
(175)
` ˘
pqp ´ pqqf “ ´ıx∇f ` ı1f ` ıx∇f “ ı1f f P SpRd q .
So the CCRs are satisfied on the core of q and p.
67
Remark: Notice, that we could as easily choose a different representation in which
U pαq would act by transition and V pβq would act by multiplication. In this case, we
would obtain the Fourier transform of the above choice. Notice, that this also implies
that the definition of position operator as a multiplication operator is ambiguous. It
really depends on the Hilbert space we choose. In fact, because of this symmetry
between representations, it is hardly possible to distinguish between position and
momentum operator at all. The only distinguishing feature will be the dynamics of
the system, so we can realize which is which, by looking at their time evolution.
ż
ωψ pAq “ ψpxqpAψqpxq dλpxq. (176)
Rd
where |ψ|2 pxq defines the probability distribution finding the particle at position x. This
interpretation is possible for the algebra generated by A. In general, however, it is not
possible to find a state that has a probabilistic interpretation for all observables. This
distinguishes quantum mechanics from the classical case.
The next step is be to define the dynamics of the particle by defining the evolution
operator. This, however, is beyond the scope of this lecture.
68
References
[1] R. V. Kadison and J. R. Ringrose. Fundamentals of the theory of operator algebras
Volume I. Vol. 1. American Mathematical Soc., 1997.
[2] R. V. Kadison and J. R. Ringrose. Fundamentals of the theory of operator algebras
Volume III. Vol. 3. American Mathematical Soc., 1997.
[3] M. Reed. Methods of modern mathematical physics I: Functional analysis. London:
Academic Press, 1980.
[4] S. Sakai. C*-algebras and W*-algebras. Heidelberg, Berlin: Springer Science &
Business Media, 2012.
[5] G. K. Pedersen. Analysis now. Vol. 118. New York: Springer Science & Business
Media, 2012.
[6] F. Strocchi. An introduction to the mathematical structure of quantum mechanics:
a short course for mathematicians. Vol. 28. Singapore: World Scientific, 2008.
[7] G. G. Emch. Algebraic methods in statistical mechanics and quantum field theory. In-
terscience monographs and texts in physics and astronomy. Rochester, New York:
Wiley-Interscience, 2009.
[8] D. Ruelle. Statistical mechanics: Rigorous results. 2nd. New York: W. A. Benjamin
Inc., 1974.
[9] O. Bratteli and D. W. Robinson. Operator Algebras and Quantum Statistical Me-
chanics 1: C*-and W*-Algebras. Symmetry Groups. Decomposition of States. Vol. 1.
Berlin, Heidelberg: Springer Science & Business Media, 2012.
[10] O. Bratteli and D. W. Robinson. Operator Algebras and Quantum Statistical Me-
chanics 2: Quantum states and Models in Quantum Statistical Mechanics. Vol. 2.
Berlin, Heidelberg: Springer Science & Business Media, 2012.
[11] H. Hanche-Olsen and E. Størmer. Jordan operator algebras. Vol. 21. London: Pit-
man Advanced Publishing Program, 1984.
[12] W. Heisenberg. The physical principles of the quantum theory. New York: Dover
Publications, INC., 1949.
[13] W. Heisenberg. “Über quantentheoretische Umdeutung kinematischer und mech-
anischer Beziehungen.” In: Zeitschrift für Physik 33.1 (1925), pp. 879–893.
[14] W. Heisenberg. “Über den anschaulichen Inhalt der quantentheoretischen Kine-
matik und Mechanik”. In: Zeitschrift für Physik 43.3 (1927), pp. 172–198.
[15] J. v. Neumann. “Die eindeutigkeit der Schrödingerschen operatoren”. In: Mathe-
matische Annalen 104.1 (1931), pp. 570–578.
69
[16] M. Born and P. Jordan. “Zur quantenmechanik”. In: Zeitschrift für Physik 34.1
(1925), pp. 858–888.
[17] W. Rossmann. Lie groups: an introduction through linear groups. Vol. 5. Oxford:
Oxford University Press on Demand, 2006.
70