This paper extends the tools of C*-algebraic strict quantization toward analyzing the classical limits of unbounded quantities in quantum theories. We introduce the approach first in the simple case of finite systems. Then, we apply this approach to analyze the classical limits of unbounded quantities in bosonic quantum field theories, with particular attention to number operators and Hamiltonians. The methods take classical limits in a representation-independent manner and so allow one to compare quantities appearing in inequivalent Fock space representations.

Powerful C*-algebraic tools have been developed in the last few decades for analyzing the classical limits of quantum theories. These tools form the theory of strict quantization1–6 (in contrast to formal quantization,7,8 which is used in perturbative quantum field theory). Working with C*-algebras provides a rich structure in which to construct quantum theories and their classical limits as well as provide physical interpretations. However, it is sometimes said that one cannot use C*-algebras to model all of the systems of physical interest. In quantum field theory, for example, researchers often employ more flexible types of *-algebras.9–11 One reason is that C*-algebras do not allow one to capture unbounded quantities such as the field operators, field momentum operators, or number operators associated with such systems. Yet, recent developments in the theory of unbounded operator algebras have made precise the relationships between algebras of bounded and unbounded quantities.12 Certain algebras of unbounded quantities can be understood as completions, in a relevant topology, of C*-algebras. In this paper, we leverage this fact to make some first steps toward understanding the classical limits of unbounded quantities starting from the framework of strict quantization. Our central contribution is to develop tools for taking the classical limits of unbounded quantities and to illustrate these tools in free bosonic quantum field theories by analyzing classical limits of number operators.

Others have analyzed unbounded quantities in C*-algebraic terms by working in specified Hilbert space representations. This allows one to consider unbounded operators affiliated with a represented C*-algebra. This technique is useful for many purposes, but its dependence on a Hilbert space representation has some drawbacks for systems whose kinematical C*-algebras have unitarily inequivalent representations, including quantum field theories.13 By contrast, the methods we develop in this paper work directly at the level of the abstract algebras, thus providing a framework that one can use to simultaneously compare even unbounded quantities that appear in inequivalent representations.

The plan of this paper is as follows. In Sec. II, we provide background on strict quantization. In Sec. III, we develop tools for taking the classical limits of unbounded quantities in a system with finitely many degrees of freedom—e.g., an n-particle system. We use this simpler example to outline key features of a strict quantization that allow one to extend it to unbounded quantities. This serves as a jumping off point for the generalization of these methods in Sec. IV to linear bosonic field theories with infinitely many degrees of freedom. In Sec. V, we focus specifically on the Klein–Gordon field and establish the classical limits of number operators associated with inequivalent representations of the kinematical C*-algebra. In Sec. VI, we apply these methods to the free Maxwell field. We conclude in Sec. VII with some discussion.

Much previous work on the classical limit precedes the approach of the present paper. For example, Hepp14 provided a semi-classical analysis of bosonic field theories in a particular Fock space representation. Recent work of Ammari et al.15 provides a different framework from the current investigation that also allows one to discuss coherent states in bosonic field theories, which are known2,16 to be closely related to the Berezin quantization map employed in the current paper. Falconi17 extended that previous work in a representation-independent manner, although the Wigner measures used in that work are associated with the alternative to Berezin quantization called Weyl quantization; since Weyl quantization is not positive, the corresponding Wigner measures are, in general, not positive.2 We will not provide a comprehensive comparison of different approaches to the classical limit here. We simply note that our work falls in the tradition of strict quantization, which Feintzeig18 argues provides an appropriate physical interpretation based on uniform approximations of observables. We will also aim to obtain uniform approximations in the classical limit here, even though we rely on pointwise approximations to generate unbounded observables.

A strict quantization provides the mathematical framework for analyzing classical limits of states and quantities within a C*-algebraic setting by giving one a notion of the limit of a family of C*-algebras. For background and examples, see the work of Rieffel4–6 and Landsman.1–3,19–22

Definition 1.

A strict quantization consists in a family of C*-algebras {A}[0,1] and a family of quantization maps {Q:PA}[0,1], each of whose domain PA0 is a Poisson algebra, with Q0 being the embedding map. We require further that Q[P] is norm dense in A for each ∈ [0, 1] and that the following conditions are satisfied for all A,BP:

  • (Dirac’s condition) lim0i[Q(A),Q(B)]Q({A,B})=0;

  • (von Neumann’s condition) lim0Q(A)Q(B)Q(AB)=0;

  • (Rieffel’s condition) the map Q(A) is continuous.

A strict quantization determines the structure of a continuous field of C*-algebras in which all sections of the form [Q(A)] for AP are continuous (see Theorem II.1.2.4, p. 111 in Ref. 2). Thus, a strict quantization defines the classical limits of quantities and states as follows. The classical limit of a family of quantities {Q(A)}[0,1] is understood to be the classical quantity AP. A family of states {ωS(A)}[0,1] is called a continuous field of states when the map ω(Q(A)) is continuous for each AP. The classical limit of a continuous field of states is understood to be the classical state ω0.

An illustrative example of strict quantization is the quantization of the Weyl algebra, which we review now and use later on. First, we define the Weyl algebra itself. Let E be a vector space of test functions with a symplectic form σ. In Sec. III, we will focus on the case where E=R2n and σ is the standard symplectic form, but in Sec. IV, we will deal with the case where E is infinite-dimensional, so we proceed here with some generality. The Weyl algebra W(E,σ) is generated by elements W(F) for each F, GE with

W(F)W(G)ei2σ(F,G)W(F+G)W(F)*W(F).

The elements of the form W(F) are linearly independent, and we denote the their linear span as Δ(E, ℏσ). There is a unique C*-norm on Δ(E, ℏσ) called the minimal regular norm.23,24 We define W(E,σ) as the completion of Δ(E, ℏσ) with respect to this norm.

The commutative algebra W(E,0) is *-isomorphic to the algebra AP(E′) of σ(E′, E)-continuous almost periodic functions on some topological dual E′ to E, when E is given a vector space topology (see p. 2902 of Ref. 23). Thus, the algebra of classical quantities can be interpreted in a natural way as an algebra of functions on a phase space E′. Furthermore, the *-algebra Δ(E, 0) carries a Poisson bracket defined as the extension of

{W0(F),W0(G)}σ(F,G)W0(F+G)

(see p. 334 of Ref. 25 or Eq. (2.15), p. 11 of Ref. 26) and so can serve as the domain of a quantization map. We note that in the special case where E=R2n, the phase space is the dual E=R2n and so W(R2n,0)AP(R2n).

When E=R2n, we have further information about the algebras W(R2n,σ) for > 0. These algebras have a familiar Hilbert space representation on the Hilbert space L2(Rn), which we denote πS:W(E,σ)B(L2(Rn)) and call the Schrödinger representation

(πS(W(a,b))ψ)(x)eiab2eibxψ(x+a)

for all a,bRn. Since the families tπS(W(ta,tb)) are weak operator continuous, Stone’s theorem (see p. 264 of Ref. 27) implies that these one-parameter unitary groups have self-adjoint generators. These generators are unbounded operators, corresponding to the standard position and momentum operators for n particles, and so this Hilbert space representation reproduces the ordinary formulation of quantum mechanics.

To define a strict quantization, we work with the C*-algebras AW(E,σ) for each ∈ [0, 1], where A0=W(E,0) contains PΔ(E,0) as a dense Poisson subalgebra. We define the Weyl quantization maps QW:Δ(E,0)W(E,σ) as the linear extension of

QW(W0(F))W(F).

Binz et al.25 showed that this structure indeed forms a strict quantization. Thus, this structure allows one to analyze classical limits of states and quantities in the Weyl algebra.

If one knows that quantization maps Q:PA not only satisfy the conditions (i)–(iii) of a strict quantization but also furthermore are continuous in a locally convex topology, then one can continuously extend these maps to the completions of the respective algebras in that topology. Recent work on algebras of unbounded operators12,28–31 shows that such a completion of a C*-algebra A, which we will, in general, denote by Ã, will be at least a partial *-algebra containing unbounded operators with some discernible structure.

For what follows, we will not need the details of the rich structure theory that has been developed for algebras of unbounded operators.32–34 Instead, the issue we encounter in applying these ideas in quantization is that quantization maps may fail to be continuous, and continuity is required to guarantee a unique extension of a quantization map to a completion. For example, the Weyl quantization maps defined in Sec. II fail to be continuous in the norm, and hence weak, topologies. This implies that one cannot continuously extend the Weyl quantization maps to the completion of the Weyl algebra. This is unfortunate because the Weyl quantization maps have some nice properties; they can be defined with the minimal algebraic structure of the Weyl algebra even on an infinite dimensional phase space. However, another quantization prescription called Berezin quantization is known to be continuous in the norm, and hence weak, topologies. Berezin quantization is well-defined for systems with finitely many degrees of freedom with phase space R2n, but the standard definition involves phase space integrals that are not, in general, well-defined when E is infinite-dimensional. Our goal in this section is thus to put Berezin quantization into a minimal algebraic form so that it can be applied even to systems whose phase space is infinite-dimensional. As we proceed, we will use the simplified example of a system with finitely many degrees of freedom to illustrate the basic concepts of our approach to dealing with classical limits of unbounded operators.

We begin by defining the Berezin quantization maps for a system with phase space R2n. We will work with the algebras AK(L2(Rn)) of compact operators on L2(Rn) for each ∈ (0, 1] and the algebra A0C0(R2n) of continuous functions vanishing at infinity on R2n, the latter of which contains the dense Poisson subalgebra PCc(R2n) of smooth, compactly supported functions. The Berezin quantization maps involve integrals over phase space of certain functions of coherent states, but since these integrals are not, in general, meaningful on infinite dimensional phase spaces, we will seek to put the quantization maps in a different form. A coherent state for (p,q)R2n is a vector ψ(p,q)L2(Rn) of the form

ψ(p,q)(x)1(π)n/4expipq2+ipx(xq)22.

Here and in what follows, x2 denotes the dot product x ⋅ x for any xRm. The Berezin quantization maps QB:Cc(R2n)K(L2(Rn)) are then defined for each fCc(R2n) by

(QB(f)ψ)(x)1(2π)nR2nf(p,q)ψ(p,q)(x)ψ(p,q),ψdpdq

for each ψL2(Rn), where ⟨⋅, ⋅⟩ is the L2 inner product.

It is known that this quantization map is positive, which implies that it is continuous in the norm (see Proposition 1.3.7, p. 47 of Ref. 2), and hence weak, topologies. First, this entails that QB extends continuously to a map C0(R2n)K(L2(Rn)), which we will denote by the same symbol. Second, this implies that QB extends continuously to the weak completions of the domain and range, which we now denote Q̃B:C0̃(R2n)K̃(L2(Rn)). The algebra C̃0(R2n) contains many unbounded and even discontinuous functions (see Example 4.1, p. 371 of Ref. 28), but for our purposes, we note that it at least contains as a subalgebra the algebra of all continuous functions C(R2n)C̃0(R2n) and so contains unbounded functions Φ0(a,b):R2nC for fixed a,bRn of the form

Φ0(a,b)(p,q)ap+bq

for each p,qRn, where ⋅ denotes the usual dot product. These functions include standard classical position and momentum observables. Similarly, K̃(L2(Rn)) contains many unbounded operators (see Example 4.3, p. 372 of Ref. 28), including all operators Φ(a, b) on L2(Rn) for fixed a,bRn of the form

(Φ(a,b)ψ)(x)(iaψ)(x)+(bx)ψ(x)

acting on the dense domain of vectors ψCc(Rn)L2(Rn). Again, these operators include standard quantum position and momentum observables.

Note that AP(R2n)C(R2n)C̃0(R2n), and so one can directly compare the maps Q̃B and QW on the domain Δ(E,0)AP(R2n) [where we freely identify W(R2n,0) with AP(R2n)]. The comparison actually follows directly from the known relationship of Weyl quantization with Berezin quantization [see Eq. (2.117), p. 144 of Ref. 2] or from the representation of Berezin quantization in terms of Toeplitz operators on a Segal–Bargmann space (see p. 294 of Ref. 35). Here, we will establish the comparison in the Schrödinger representation of the Weyl algebra on L2(Rn) by direct computation. (See also Ref. 36 for a generalization related to Rieffel’s deformation.)

Define c(a,b)=e4(a2+b2) for all (a,b)R2n. Let cQW:Δ(E,0)W(R2n,σ) be the linear extension of the map defined on the generators W0(a,b)AP(R2n) by

cQW(W0(a,b))c(a,b)QW(W0(a,b)).

The following proposition establishes that cQW is equivalent to the extension of QB:

Proposition 1.

For anyfAP(R2n),πScQW(f)=Q̃B(f). In other words, the diagram inFig. 1commutes.

FIG. 1.

Commutative diagram for Proposition 1.

FIG. 1.

Commutative diagram for Proposition 1.

Close modal

Proof.

It suffices to show that the identity holds on the generators W0(a, b) ∈ Δ(E, 0) for arbitrary a,bRn.

To show this, we first note that the Fourier inversion theorem implies that for any ψL2(Rn),
ψ(x+a)expiab2a24=RnRnexp(2πi(ay)ξ)ψ(x+y)expiby2y24dydξ.
Setting p = 2πℏξ gives the equation
ψ(x+a)expiab2a24=1(2π)nRnRnexpi(ay)pψ(x+y)expiby2y24dydp=1(2π)nRnRnψ(x+y)expiap+iby2ipyy24dydp.
This implies that
(Q̃B(W0(a,b))ψ)(x)=1(π)n/2(2π)nRnRnRnψ(y)expiap+ibq+ipx(xq)22ipy(yq)22dqdydp=1(π)n/2(2π)nRnRnRnψ(y)expq2+ib+x+yq+iap+ipxx22ipyy22dqdydp=1(2π)nRnRnψ(y)exp4ib+x+y2+iap+ipxx22ipyy22dydp=1(2π)nRnRnψ(y)expb24+x24+y24+ibx2+iby2+xy2+iap+ipxx22ipyy22dydp=1(2π)nRnRnψ(y)expb24+ibx2+iby2+xy2+iap+ipxx24ipyy24dydp=1(2π)nRnRnψ(y)expb24+ibx+iap+ib(yx)2ip(yx)(yx)24dydp=expb24+ibx1(2π)nRnRnψ(x+y)expiap+iby2ipyy24dydp=expb24+ibxexpiab2a24ψ(x+a)=c(a,b)expiab2+ibxψ(x+a)=(πS(cQW(W0(a,b)))ψ)(x),
which is the desired result.□

The scalars c form what Honegger and Rieckers37 call quantization factors, satisfying

  • c(a,b)R+ for all ∈ [0, 1] and a,bRn;

  • c(0, 0) = e0 = 1 and c0(a, b) = e0 = 1 for all ∈ [0, 1] and a,bRn; and

  • c(a,b)=e4(a,b) is continuous for all a,bRn.

This implies (by Theorem 4.4, p. 129 of Ref. 37) that the maps cQW likewise define a strict quantization. Thus, we can use the maps cQW to provide a definition of the Berezin quantization on the minimal algebraic structure of the Weyl algebra. Since Berezin quantization is positive, and hence continuous, we can extend these maps to unbounded operators defined from the Weyl algebra.

Our goal is to use the quantization maps cQW to analyze the classical limits of unbounded operators such as Φ(a, b). To do so, we note that these operators can be constructed from the unitary generators W(a, b) of the Weyl algebra by the formal relation

Φ(a,b)ilimt0W(ta,tb)It.
(1)

This relation holds strictly in the Schrödinger representation when the limit is understood in the weak operator topology on B(L2(Rn)). However, the limit does not, in general, converge in the abstract weak topology on W(R2n,σ). In the service of our goal of analyzing quantization in a representation manner, we seek a different abstract algebra with a natural topology in which these limits converge. Then, we will be able to use Eq. (1) as a definition of Φ(a, b) solely in terms of abstract algebraic structure.

To construct such an algebra, we will form the quotient algebra by a certain two-sided ideal. Our ultimate goal is to find an algebra that allows only for states whose expectation values of Eq. (1) to converge. It is known38,39 that one can limit the collection of states of an algebra if one chooses to quotient by an ideal that is the annihilator of the set of states one wants to focus on. More precisely, given a C*-algebra B and a collection of functionals VB*, under certain conditions on V, one can construct a new C*-algebra A whose dual space contains only the functionals in V, i.e., A*V. To do so, first let N(V) denote the annihilator of V in B. If N(V) is a closed, two-sided ideal in B, then setting A=B/N(V) produces a C*-algebra with the desired dual space.

This is relevant to the current circumstance if we focus on the states on the Weyl algebra for which the expectation values of Eq. (1) converge. To that end, we focus on the so-called regular states and define

V{ωW(R2n,σ)*|tω(W(ta,tb)) is continuous for every a,bRn}.

However, in this case, N(V) is not a closed, two-sided ideal in W(R2n,σ) because the latter algebra is simple. Hence, we move to the bidual W(R2n,σ)** and consider the weak* closure V¯W(R2n,σ)*** of V, understood now as the regular functionals on the bidual. It follows that the annihilator N(V¯) in W(R2n,σ)** is now a closed, two-sided ideal. Hence, we can complete the construction by defining a quotient C*-algebra AW(R2n,σ)**/N(V¯).

It follows that A*V¯. Moreover, the algebra A0 is *-isomorphic to the algebra BR(R2n) of bounded universally Radon measurable functions, and the algebras A are *-isomorphic to B(L2(Rn)) for each > 0.39 Thus, we have canonical projection (quotient) maps p0:AP(R2n)**BR(R2n) and p:W(R2n,σ)**B(L2(Rn), and even further, we have BR(R2n)C0(R2n)** and B(L2(Rn))K(L2(Rn))** so that both algebras are W*-algebras carrying natural weak* topologies. The families tp(W(ta, tb)) are weak* continuous in A for all ∈ [0, 1], so the limit in Eq. (1) is well-defined in the weak* topology.

We are now in a position to consider the functions Φ0(a, b) in the domain of our quantization maps. To do so, we continuously extend cQW in the weak topology to a map cQ̃W:AP(R2n)**W(R2n,σ)**. We have the following corollary of Proposition 1.

Corollary 1.

For anyfAP(R2n)**,pcQ̃W(f)=Q̃Bp0(f). In other words, the diagram inFig. 2commutes.

This informs us that the map pcQ̃W, which we emphasize can be defined in terms of abstract algebraic structure, is a positive quantization map equivalent to Berezin quantization on R2n. Thus, pcQ̃W extends continuously to the entire map Q̃B:C̃0(R2n)K̃(L2(Rn)). We can understand Φ0(a, b) to be defined in the domain C̃0(R2n) and Φ(a, b) to be defined in the range K̃(L2(Rn)) both via Eq. (1), where the limits are in the abstract weak* topologies.

Finally, we note that the conditions of a strict quantization extend to the unbounded operators Φ(a, b). The results mentioned now are familiar consequences of Eq. (1) and the algebraic relations in the Weyl algebra. We establish them explicitly in the more general setting of Sec. IV, but we state them here already. First, the quantization map assigns Q̃B(Φ0(a,b))=Φ(a,b). Second, the canonical commutation relations are satisfied,
[Q̃B(Φ0(a,b)),Q̃B(Φ0(a,b))]=iσ((a,b),(a,b))I.
This implies that Dirac’s condition is satisfied in the form
lim0i[Q̃B(Φ0(a,b)),Q̃B(Φ0(a,b))]Q̃B({Φ0(a,b),Φ0(a,b)})=0,
with the use of the standard Poisson bracket on R2n. Furthermore, von Neumann’s condition is satisfied in the form
lim0Q̃B(Φ0(a,b))Q̃B(Φ0(a,b))Q̃B(Φ0(a,b)Φ0(a,b))=0.
Thus, there is a strong sense in which the functions Φ0(a, b) can be understood as the classical limits of the operators Φ(a, b).

FIG. 2.

Commutative diagram for Corollary 1.

FIG. 2.

Commutative diagram for Corollary 1.

Close modal

Suppose now that E is an infinite dimensional vector space with a symplectic form σ. This is the case when E is the test function space for any free Bosonic field theory whose phase space E′ is a linear space. Although the integral formulas defining Berezin quantization in Sec. III A are no longer meaningful in this context, we proceed to construct an analogous positive quantization, which can likewise be extended to unbounded operators.

We start with the Weyl quantization maps QW, which are well-defined even in the infinite-dimensional setting, and we aim to define quantization factors in the spirit of Sec. III A. We require a norm on E, which may be determined as follows. Suppose that we are given a complex structureJ : EE compatible with σ—that is, a linear map satisfying

  • σ(JF, JG) = σ(F, G);

  • σ(F, JF) ≥ 0; and

  • J2 = −I

for all F, GE. In general, there is no such unique complex structure; we will see concrete examples below. A complex structure can be used to define a complex inner product

αJ(F,G)σ(F,JG)+iσ(F,G)

for all F, GE. This inner product αJ allows us to define quantization factors cJ:ER+ by cJ(F)e4αJ(F,F). These quantization factors satisfy the same conditions (a)–(c) of Sec. III A. Now, in analogy with Sec. III A, we define new quantization maps QJ:Δ(E,0)W(E,σ) by the linear extension of

QJ(W0(F))cJ(F)QW(W0(F)).

For any choice of complex structure J, this defines a strict quantization equivalent to QW in the sense that (see Theorem 4.6, p. 131 of Ref. 37)

lim0QW(A)QJ(A)=0

for all A ∈ Δ(E, 0). It follows that the strict quantizations defined for different choices of complex structure J and J′ are also all equivalent in this same sense as

lim0QJ(A)QJ(A)=0

for all A ∈ Δ(E, 0).

One can show that the quantization maps QJ possess some of the same virtues as the Berezin quantization maps of Sec. III A.

Proposition 2.

IfJis a complex structure compatible with the symplectic formσ, then the mapQJ:Δ(E,0)W(E,σ)is positive.

Proof.
Suppose C ∈ Δ(E, 0) is a positive element. Then, C = A*A for some A = ∑kzkW0(Fk) ∈ Δ(E, 0). We have
QJ(A*A)=j,kz¯jzkexp4αJ(Fj+Fk,Fj+Fk)W(FkFj)=j,kz¯jzkexp4αJ(Fj,Fj)+αJ(Fk,Fk)+2σ(Fj,JFk)i2σ(Fj,Fk)W(Fj)W(Fk).
Letting yk=e4αJ(Fk,Fk)zk, it follows that
QJ(A*A)=jky¯jykexp2σ(Fj,JFk)+iσ(Fj,Fk)W(Fj)*W(Fk)=jky¯jykexp2αJ(Fj,Fk)W(Fj)*W(Fk).
Since αJ is a complex inner product, the matrix aj,k : = αJ(Fj, Fk) is positive, and moreover, since entrywise exponentiation preserves positivity, the matrix bj,k ≔ exp(aj,k) is also positive. It then follows from a generalization of the Schur product theorem due to Ref. 40 (Proposition 1.3) that QJ(A*A) is a positive element in W(E,σ).□

The positivity of QJ implies its continuity in the norm and weak topologies (see Proposition 1.3.7 of Ref. 2, p. 47), which means that it can be continuously extended to the completions of its domain and range. As in Sec. III A, we want to use these extended quantization maps to analyze field operators of the form

Φ(F)ilimt0W(tF)It.
(2)

However, these limits again do not converge in the weak topology. So we must perform the construction of Sec. III B to arrive at a new algebra allowing for this definition.

To construct such an algebra, we again quotient out by a certain two-sided ideal given by the annihilator of a desired set of states. We again focus on the states for which the expectation values of Eq. (2) converge by defining the set of regular states as

V{ωW(E,σ)*|tω(tF) is continuous for every FE}.

Just as before, N(V) is not a closed, two-sided ideal because W(E,σ) is simple. Instead, we use the strategy of Sec. III B by passing to the bidual W(E,σ)** and letting V¯ be the weak* closure of V in W(E,σ)***. Then, N(V¯) is a closed, two sided ideal, so we can define a C*-algebra AW(E,σ)**/N(V¯) exactly as before.

However, since E is now infinite-dimensional and so fails to be locally compact, the structure of these algebras A is not as tractable, and we have much less information than in Sec. III B. Still, we can show that the algebras A are W*-algebras with an appropriate weak* topology.

Proposition 3.

The algebrasAare W*-algebras with preduals given by(A)*=V.

Proof.

First, let πU denote the universal representation of W(E,σ). We will consider the direct sum representation πωIπω for I=VS(W(E,σ)), where πω is the GNS representation for the state ω and S(W(E,σ)) denotes the state space of the Weyl algebra. It follows from Theorem 10.1.12 of Ref. 41 (p. 719) that there is a projection P in the center of πU(W(E,σ)¯, where the closure is in the weak operator topology, such that πU(W(E,σ)¯P is *-isomorphic to π(W(E,σ)¯, the latter of which is *-isomorphic to A. By Proposition 5.5.6 of Ref. 41 (p. 335), the algebra πU(W(E,σ)¯P is a W*-algebra, which implies that A is a W*-algebra. Moreover, by Proposition 5 of Ref. 42 (p. 15), it follows that (A)*=V.□

This implies that QJ extends continuously to a map whose codomain is the weak* completion Ã, which we now denote Q̃J:Ã0Ã. The field operators are well-defined in these completed algebras via Eq. (2) with the limit now understood in the weak* topology. Now we can use the maps Q̃J to analyze the classical limits of unbounded field operators.

First, we note that the familiar facts about the field operators Φ(F) follow from what has been said so far. We present proofs here to emphasize the fact that these statements can be both expressed and derived in the bare algebraic setting we have outlined.

Lemma 1.

For allFE, Φ(F) is self-adjoint.

Proof.
For any FE, we have
(Φ(F))*=ilimt0W(tF)It*=ilimt0W(tF)It=ilims0W(sF)Is=Φ(F).
In the third line, we make the replacement s = −t.□

Lemma 2.

For allFE,Q̃J(Φ0(F))=Φ(F). In other words, the diagram inFig. 3commutes.

FIG. 3.

Commutative diagram for Lemma 2.

FIG. 3.

Commutative diagram for Lemma 2.

Close modal

Proof.
For any FE, we have
Q̃J(Φ0(F))=Q̃Jilimt0W0(tF)It=ilimt0e4αJ(tF,tF)W(tF)It=ilimt0e4αJ(tF,tF)W(tF)W(tF)tilimt0W(tF)It=ilimt0(e4t2αJ(F,F)1)W(tF)t+Φ(F)=ilimt0e4t2αJ(F,F)1tlimt0W(tF)+Φ(F)=i(0)(I)+Φ(F)=Φ(F).

Lemma 3.

For allF, GEand allnN,(F), Φ(G)n] = inℏσ(F, G(G)n−1.

Proof.
First, we compute
[Φ(F),Φ(G)]=ilims0W(sF)Isilimt0W(tG)Itilimt0W(tG)Itilims0W(sF)Is=lims0limt01st(W(tG)I)(W(sF)I)(W(sF)I)(W(tG)I)=lims0limt01stei2σ(tG,sF)ei2σ(sF,tG)W(sF+tG)=lims0limt0ei2stσ(G,F)ei2stσ(F,G)stlims0limt0W(sF+tG)=iσ(F,G)I.
Next, we proceed by induction. Suppose [Φ(F),Φ(G)k]=ikσ(f,g)(Φ(G))k1 for some kN. Then,
Φ(F)Φ(G)k+1=(Φ(F)Φ(G)k)Φ(G)=ikσ(F,G)Φ(G)k1+Φ(G)kΦ(F)Φ(G)=ikσ(F,G)Φ(G)k+Φ(G)kiσ(F,G)I+Φ(G)Φ(F)=i(k+1)σ(F,G)Φ(G)k+Φ(G)k+1Φ(F),
which implies [Φ(F), Φ(G)k+1] = i(k + 1)ℏσ(F, G(G)k.□

Lemma 4.

For allFE,Q̃J(Φ0(F)2)=Φ(F)2+2αJ(F,F)I.

Proof.
First, we compute
Q̃J(Φ0(F)2)=Q̃J(i)2lims0limt0(W0(sF)I)(W0(tF)I)st=Q̃Jlims0limt0W0((s+t)F)W0(sF)W0(tF)+Ist=lims0limt0e4(s+t)2αJ(F,F)W((s+t)F)e4s2αJ(F,F)W(sF)e4t2αJ(F,F)W(tF)+Ist.
Compare this to the value
Q̃J(Φ0(F))2=(i)2limt0QJ(W0(tF))It2=lims0limt0e4s2αJ(F,F)W(sF)Ie4t2αJ(F,F)W(tF)Ist=lims0limt0e4(s2+t2)αJ(F,F)W((s+t)F)e4s2αJ(F,F)W(sF)e4t2αJ(F,F)W(tF)+Ist.
This gives the identity
Q̃J(Φ0(F)2)Q̃(Φ0(F))2=lims0limt0e4(s+t)2αJ(F,F)e4(s2+t2)αJ(F,F)stW((s+t)F)=lims0limt0e4(s2+t2)αJ(F,F)e4(s+t)2αJ(F,F)stlims0limt0W((s+t)F)=2αJ(F,F)I.
Thus, Q̃J(Φ0(F)2)=Q̃J(Φ0(F))2+2αJ(F,F)I=Φ(F)2+2αJ(F,F)I.□

We would like to extend the conditions of a strict quantization to even unbounded operators such as Φ(F). However, since we have extended the quantization map in the weak topology, the resulting notion of approximation in the classical limit is significantly weaker than the norm approximations in a strict quantization. We do at least have a notion of approximation pointwise on each state, as follows. Fix some choice of H ∈ [0, 1] and an arbitrary functional ωHVH. We construct the “constant” section of linear functionals {ωV}[0,1] through the point ωH as the continuous extension of

ω:QJ(A)ωH(QHJ(A))C

for each ∈ [0, 1] and each A ∈ Δ(E, 0). Then, for any A,BÃ0 and any ϵ > 0, there is an ′ ∈ (0, 1] such that for all < ′,

ωQ̃J(A)Q̃J(B)Q̃J(AB)<ϵ

when AB and Q̃J(A)Q̃J(B) are well-defined. This approximation is of course much weaker than one would like. However, we show next that the preliminary lemmas just stated imply that the classical limits of field operators, in particular, satisfy a stronger approximation given by Dirac’s condition and von Neumann’s condition for a strict quantization. This follows because although the field operators are unbounded and the norm is not defined on them, the relevant differences of operators are bounded and so the conditions are meaningful exactly as stated. In what follows, we understand the Poisson bracket to be defined as in Ref. 26, Eq. (2.15), p. 11; cf. the Peierls bracket as defined in Refs. 43 and 11.

Proposition 4.

For allF, GE,

  1. lim0i[Φ(F),W(G)]Q̃J({Φ0(F),W0(G)})=0.

  2. lim0Φ(F)W(G)Q̃J(Φ0(F)W0(G))=0.

  3. lim0i[Φ(F),Φ(G)]Q̃J({Φ0(F),Φ0(G)})=0.

  4. lim0Φ(F)Φ(G)Q̃J(Φ0(F)Φ0(G))=0.

Proof.
1. First, we have {Φ0(F), W0(G)} = (G, F)W0(G), which implies
Q̃J({Φ0(F),W0(G)}=iσ(G,F)e4αJ(G,G)W(G).
Furthermore,
Φ(F)W(G)=ilimt0W(tF)ItW(G)=W(G)ilimt0eiσ(G,tF)W(tF)It=W(G)ilimt0eiσ(G,tF)1tIilimt0W(tF)It=W(G)((i)(i)σ(G,F)I+Φ(F))=W(G)Φ(F)+σ(G,F)W(G),
which implies
i[Φ(F),W(G)]=iσ(G,F)W(G),
and hence,
lim0i[Φ(F),W(G)]Q̃J({Φ0(F),W0(G)})=lim0iσ(G,F)iσ(G,F)e4αJ(G,G)W(G)=lim0|iσ(G,F)(1e4αJ(G,G))|=0.
  1. We have
    Φ0(F)W0(G)=ilimt0W0(tF+G)W0(G)t
    so that
    Q̃J(Φ0(F)W0(G))=ilimt0e4αJ(tF+G,tF+G)W(tF+G)e4αJ(G,G)W(G)t=ilimt0e4αJ(tF+G,tF+G)ei2σ(tF,G)W(tF)e4αJ(G,G)ItW(G)=ilimt0e4αJ(tF+G,tF+G)ei2σ(tF,G)1tIie4αJ(G,G)limt0W(tF)ItW(G)=i2σ(F,JG)I+e4αJ(G,G)Φ(F)W(G).
    Consequently,
    lim0Φ(F)W(G)Q̃J(Φ0(F)W0(G))=lim0|i2σ(F,JG)|=0.
  2. This follows immediately from Lemma 3 together with the fact that
    {Φ0(F),Φ0(G)}=σ(F,G).
  3. We have
    Q̃J(Φ0(F)Φ0(G))=(i)2Q̃Jlimt0lims0W0(tF+sG)W0(tF)W0(sG)+Ist=(i)2limt0lims0e4αJ(tF+sG,tF+sG)W(tF+sG)e4αJ(tF,tF)W(tF)e4αJ(sG,sG)W(sG)+Ist
    and
    Q̃J(Φ0(F))Q̃J(Φ0(G))=(i)2limt0e4αJ(tF,tF)W(tF)Itlims0e4αJ(sG,sG)W(sG)Is=(i)2limt0lims01ste4(αJ(tF,tF)+αJ(sG,sG))ei2σ(tF,sG)W(tF+sG)e4αJ(tF,tF)W(tF)e4αJ(sG,sG)W(sG)+I,
    which implies
    Q̃J(Φ0(F))Q̃J(Φ0(G))Q̃J(Φ0(F)Φ0(G))=(i)2limt0lims01ste4(αJ(tF,tF)+αJ(sG,sG))ei2stσ(F,G)e2stαJ(F,G)W(tF+sG)=limt0lims0ei2stσ(F,G)e2stσ(F,JG)1stI=2σ(F,JG),
    and hence,
    lim0Φ(F)Φ(G)Q̃J(Φ0(F)Φ0(G))=lim0|2σ(F,JG)|=0.

This establishes a strong sense in which Φ0(F) is the classical limit of Φ(F).

Suppose further that we are given a complex structure J0 compatible with σ, which may be distinct from the complex structure J used to define the quantization map. In Sec. V, we will consider two such possible complex structures. It is important to note that facts about the classical limits of quantities defined in the quantum theory via a complex structure do not depend on which complex structure is used in the definition of the quantization map. We can use a complex structure to define J0-creation and J0-annihilation operators abstractly by

aJ0(F)12Φ(F)+iΦ(J0F)aJ0(F)*12Φ(F)iΦ(J0F).

Note that it follows immediately from Lemma 2 that QJ(a0J0(F))=aJ0(F) even when J and J0 are distinct. Similarly, the J0-creation and J0-annihilation operators can be used to abstractly define the J0-number operators

NJ0(F)aJ0(F)*aJ0(F).

Although QJ(N0J0(F))NJ0(F), we can still show a sense in which N0J0(F) is the classical limit of NJ0(F).

Proposition 5.

For allFE,lim0Q̃J(N0J0(F))NJ0(F)=0.

Proof.
Note that
NJ0(F)=12Φ(F)2+Φ(J0F)2+i[Φ(F),Φ(J0F)]
and
Q̃J(N0J0(F))=12Q̃JΦ0(F)2+Φ0(J0F)2=12Φ(F)2+Φ(J0F)2+2(αJ(F,F)+αJ(J0F,J0F))=12Φ(F)2+Φ(J0F)2+αJ(F,F),
so since [Φ(F),Φ(J0F)]=iαJ0(F,F), it follows that
lim0Q̃J(N0J0(F))NJ0(F)=lim0|2(αJ(F,F)+αJ0(F,F))|=0.

Furthermore, one can show that Dirac’s condition and von Neumann’s condition hold for some combinations of creation (or annihilation) operators and number operators.

Proposition 6.

For anyF, GE,

  1. lim0i[aJ0(F),W(G)]Q̃J({a0J0(F),W0(G)})=0.

  2. lim0aJ0(F)W(G)Q̃J(a0J0(F)W0(G))=0.

  3. lim0i[aJ0(F),Φ(G)]Q̃J({a0J0(F),Φ0(G)})=0.

  4. lim0aJ0(F)Φ(G)Q̃J(a0J0(F)Φ0(G))=0.

  5. lim0iaJ0(F)*,aJ0(G)Q̃Ja0J0(F)*,a0J0(G)=0.

  6. lim0aJ0(F)*aJ0(G)Q̃Ja0J0(F)*a0J0(G)=0.

Proof.

This follows immediately from Proposition 4 along with the linearity of the creation and annihilation operators with respect to the field operators.□

Proposition 7.

For allF, GE,

  1. lim0i[NJ0(F),W(G)]Q̃J({N0J0(F),W0(G)})=0.

  2. lim0i[NJ0(F),Φ(G)]Q̃J0({N0J0(F),Φ0(G)})=0.

  3. lim0i[NJ0(F),aJ0(G)]Q̃J({N0J0(F),a0J0(G)})=0.

  4. lim0i[NJ0(F),NJ0(G)]Q̃J({N0J0(F),N0J0(G)})=0.

Proof.
1. We have
{N0J0(F),W0(G)}=iσ(G,F)Φ0(F)W0(G)+σ(G,J0F)Φ0(J0F)W0(G)
and
NJ0(F),W(G)=12[Φ(F)2+Φ(J0F)2,W(G)]=12[Φ(F),W(G)]Φ(F)+Φ(F)[Φ(F),W(G)]+[Φ(J0F),W(G)]Φ(J0F)+Φ(J0F)[Φ(J0F)),W(G)]=2σ(G,F)W(G)Φ(F)+Φ(F)W(G)+2σ(G,J0F)W(G)Φ(J0F)+Φ(J0F)W(G).
Then, (2) of Proposition 4 implies the result.
  1. We have
    {N0J0(F),Φ0(G)}=σ(G,F)Φ0(F)+σ(G,J0F)Φ0(J0F)
    and
    [NJ0(F),Φ(G)]=12[Φ(F)2+Φ(J0F)2,Φ(G)]=iσ(F,G)Φ(F)+σ(J0F,G)Φ(J0F).
    Then, Lemma 2 implies the result.
  2. This follows from (2) and the linearity of aJ0(G) with respect to the fields.

  3. We have
    {N0J0(F),N0J0(G)}=σ(G,F)Φ0(F)Φ0(G)+σ(J0G,F)Φ0(F)Φ0(J0G)+σ(G,J0F)Φ0(J0F)Φ0(G)+σ(G,F)Φ0(J0F)Φ0(J0G)
    and
    [NJ0(F),NJ0(G)]=14[Φ(F)2+Φ(J0F)2,Φ(G)2+Φ(J0G)2]=i2σ(F,G)Φ(F)Φ(G)+Φ(G)Φ(F)+i2σ(F,J0G)Φ(F)Φ(J0G)+Φ(J0G)Φ(F)+i2σ(J0F,G)Φ(J0F)Φ(G)+Φ(G)Φ(J0F)+i2σ(F,G)Φ(J0F)Φ(J0G)+Φ(J0G)Φ(J0F).
    Then, (4) of Proposition 4 implies the result.□

These results are somewhat restricted. For example, it is difficult to establish an analog of von Neumann’s condition for number operators because one encounters unbounded operators in the relevant differences. Still, we take the foregoing to establish some sense in which a0J0(F), (a0J0(F))*, and N0J0(F) are classical limits of aJ0(F), (aJ0(F))*, and NJ0(F). We recognize, however, that it would be interesting to be able to strengthen the approximations involved in the classical limit for unbounded quantities.

We now analyze the classical limits of number operators and Hamiltonians in the model of a real scalar field φ on Minkowski spacetime R4 satisfying the Klein–Gordon equation

2t22φ=m2φ,

where ∇2 is the spatial Laplacian and m > 0. We work with initial data on R3, defining ECc(R3)Cc(R3) as the space of pairs of test functions with the symplectic form

σ((f1,g1),(f2,g2)):=R3f1g2f2g1

for all f1,f2,g1,g2Cc(R3). The phase space E′ will be a topological dual to E in some vector space topology such that C(R3)C(R3)E. The space E consists in pairs (π, φ) of (possibly distributional) field configurations φ and conjugate momenta πφt. We will analyze the classical limits of two classes of number operators associated with the scalar field: “Minkowski” number operators associated with an inertial observer and “Rindler” number operators associated with an accelerating observer on the right Rindler wedge.

To define the Minkowski number operators, we must specify a choice of complex structure. To do so, we define an operator μM:Cc(R3)Cc(R3) by

μMm221/2.

This operator μM is self-adjoint and a bijection (this follows, e.g., from Theorem IX.27, p. 54 of Ref. 44). We define a complex structure JM : EE by

JM(f,g)(μM1g,μMf)

for all f,gCc(R3). JM is the unique complex structure compatible with time translations with respect to the inertial time-like symmetries of Minkowski spacetime; see Refs. 45 and 46. We define the Minkowski number operators NM(F) as the number operators corresponding to this choice of complex structure, i.e., NM(F)NJM(F) for each FE. We use similar notation for aM(F). Explicitly, we have

aM(F)Φ(F)+iΦ(JMF),NM(F)(aM(F))*aM(F).

These number operators are the usual ones defined in the Fock space representation of the Weyl algebra when the inertial time-like symmetries of Minkowski spacetime are used in the frequency splitting procedure for “second quantization.”47 

The results of Sec. IV B establish a sense in which N0M(F) is the classical limit of NM(F). We now analyze the contents of N0M(F) in the classical field theory.

We can use this setup to analyze N0M(F) as a function on E′. Recall that W(E,0) is *-isomorphic to the algebra AP(E′) of σ(E′, E)-continuous almost periodic functions on E′. In this setting, given test functions (f, g) ∈ E, the classical Weyl unitaries and fields have the form

W0(f,g)(π,φ)=expiR3(πf+φg),Φ0(f,g)(π,φ)=R3(πf+φg),

for all field configurations and conjugate momenta (π,φ)C(R3)C(R3)E. This also immediately determines the form of N0M(f,g).

Proposition 8.
For any (f, g) ∈ E,
N0M(f,g)(π,φ)=12R3πf+φg2+12R3φ(μMf)π(μM1g)2
for all(π,φ)C(R3)C(R3)E.

Furthermore, we can construct the classical Minkowski total number operatorN¯0M by letting {Fk} be any αJM-orthonormal basis for E and defining

N¯0MkN0M(Fk).

The following proposition provides an explicit form for the total number operator as a real-valued function on the phase space. This establishes that the definition of the total number operator is independent of the chosen basis, which holds similarly for all total number operators and total Hamiltonians in the remainder of this paper.

Proposition 9.
N¯0M(π,φ)=12R3π(μM1π)+φ(μMφ)
for any(π,φ)Cc(R3)Cc(R3)E.

Proof.
The Pythagorean theorem for the Hilbert space completion of E with inner product αJM implies that
N¯0M(π,φ)=kN0M(Fk)(π,φ)=12k|αJM(φ,π),Fk|2=12αJM(φ,π),(φ,π)=12R3π(μM1π)+φ(μMφ).

Finally, we can construct the classical Minkowski HamiltonianH0M. Let {fk} be any orthonormal basis for L2(R3,R) and define

H0M:=kN0M(fk,0)=kN0M(0,μMfk).

Notice that we take the sum over only one test function component and that we use an orthonormal basis with respect to the L2 inner product rather than the inner product αJM. With this definition, the classical limit of the Minkowski Hamiltonian takes a familiar form as a real-valued function on phase space.

Proposition 10.
H0M(π,φ)=12R3π2+m2φ2+(φ)2
for any(π,φ)Cc(R3)Cc(R3)E.

Proof.
It follows from the Pythagorean theorem for real Hilbert spaces and the self-adjointness of μM that
H0M(π,φ)=k12R3πfk2+R3φ(μMfk)2=k12R3πfk2+R3fk(μmφ)2=12R3π2+(μMφ)2=12R3π2+φ(μM2φ)=12R3π2+m2φ2φ(2φ)=12R3π2+m2φ2+(φ)2.
The final line is implied by the divergence theorem because
div(φφ)=φ2φ+(φ)2.

This shows that the classical limit of the Minkowski total number operator is equal to the classical total energy of the Klein–Gordon field for initial data of compact support. This, of course, is the conserved quantity of the Klein–Gordon field corresponding to the inertial time-like symmetries of Minkowski spacetime.

To define the Rindler number operators, we specify a different choice of complex structure. We work on the right Rindler wedge {(t,x,y,z)R4|x>|t|}, and so we restrict attention to initial data with support in R{(x,y,z)R3|x>0}, and we restrict attention to pairs of test functions in ECc(R)Cc(R). For comparison with Ref. 46, we work with functions of the form exf, and hence, we identify each fC(R) with fexf. We proceed as in Sec. V A by first defining an operator μR:Cc(R)Cc(R) by

μRe2xm2y2z2x21/2.

Reference 46, p. 72 establishes that μR is positive and essentially self-adjoint on E. As in Sec. V A, we define a complex structure JR:EE by

JR(f,g)(μR1g,μRf)

for all f,gCc(R). JR is the unique complex structure compatible with time translation with respect to the Lorentz boost time-like symmetries of Minkowski spacetime; see Refs. 45 and 46 We define the Rindler number operators as the number operators corresponding to this choice of complex structure, i.e., NR(f,g)NJR(f,g) for each (f,g)E. We use a similar notation for aR(f,g). Explicitly, we have

aR(F)Φ(F)+iΦ(JRF),NR(F)(aR(F))*aR(F).

These number operators correspond to those in the Fock space determined by the one-particle structure for an observer in uniform acceleration, for whom Rindler coordinates on the right Rindler wedge form a natural reference frame for dynamics, as described in Sec. 4 of Ref. 46 (see also Ref. 48).

The results of Sec. IV B establish a sense in which N0R(F) is the classical limit of NR(F). We now analyze the contents of N0R(F) in the classical field theory. That is, we again analyze N0R(F) as a function on E in the representation of W(E,0) as AP(E). As before, the representation immediately determines the form of N0R(f,g).

Proposition 11.
For any(f,g)E,
N0R(f,g)(π,φ)=12Rπf+φg2+12Rφ(μRf)π(μR1g)2
for all(π,φ)C(R)C(R)E.

Furthermore, we can construct the classical Rindler total number operatorN¯0R by letting {Fk} be any αJR-orthonormal basis for E and defining

N¯0RkN0R(Fk).

The following proposition provides an explicit form for the Rindler total number operator as a real-valued function on phase space:

Proposition 12.
N¯0R(π,φ)=12Rπ(μR1π)+φ(μRφ)
for all(π,φ)Cc(R)Cc(R)E.

Proof.

This follows from an analogous calculation to that in the Proof of Proposition 9.□

Finally, we can construct the classical Rindler HamiltonianH0R. Let {fk} be any orthonormal basis for L2(R,R) and define

H0RkN0R(fk,0)=kN0R(0,μRfk).

With this definition, we have the following explicit form of H0R as a real-valued function on phase space:

Proposition 13.
H0R(π,φ)=12R(π)2+φ(μR2φ)
for all(π,φ)Cc(R)Cc(R)E.

Proof.

This follows from an analogous calculation to that in the Proof of Proposition 10 using the self-adjointness of μR.□

This expression is the Rindler energy, which is the conserved quantity of the Klein–Gordon field associated with the time-like Lorentz boost symmetries of R, i.e., time translations in Rindler coordinates.46 Thus, the previous proposition shows that the classical limit of the Rindler total number operator is the Rindler energy.

In this section, we analyze the classical limit of the Minkowski number operator and Hamiltonian for an electromagnetic field on Minkowski spacetime satisfying the source-free Maxwell equations. As before, we work with initial data on a surface R3, on which we assume the electromagnetic field to be decomposed into an electric field vector E with components Ek (for k = 1, 2, 3) in some fixed coordinate system and a magnetic (co)vector potential A with components Aj (for j = 1, 2, 3) in the Coulomb gauge [satisfying div(A) = 0]. In this formulation, the source-free Maxwell equations take the form

div(E)=0,2t22Aj=0(for j=1,2,3).

Thus, each component Aj satisfies the mass zero Klein–Gordon equation. We take the test function space to be

V{(f,g)Tc0,1(R3)Tc1,0(R3)|div(f)=div(g)=0},

where Tc0,1(R3) is the space of smooth, compactly supported covector fields f and Tc1,0(R3) is the space of smooth, compactly supported vector fields g. We define the symplectic form on V by

σ((f,g),(f̃,g̃))R3kfkg̃kf̃kgk.

The phase space V′ will be the topological dual to V in some vector space topology, consisting of pairs (E, A) of (possibly distributional) field configurations A and conjugate momenta Ekδkj(At)j (the Euclidean metric tensor δkj is defined by δkj = 1 if k = j and 0 otherwise).

We will need the following lemma, which establishes that when E and A are smooth field configurations, they can always be chosen to be divergence free:

Lemma 5.
SupposeρVhas the form
ρ(f,g)=R3kfkEk+gkAk
for every (f, g) ∈ Vfor some smooth fields(E,A)Tc1,0(R3)Tc0,1(R3)V. Then, there is a unique pair(Ê,Â)Tc1,0(R3)Tc0,1(R3)Vwithdiv(Ê)=div(Â)=0such that
ρ(f,g)=R3kfkÊk+gkÂk
for all (f, g) ∈ V.

Proof.
The fundamental theorem of vector calculus implies that E and A can be decomposed uniquely into curl-free and divergence-free fields
Ek=Êk+δkjjϕE,Ak=Âk+kϕA,
where div(Ê)=div(Â)=0 and ϕE, ϕA are smooth scalar fields.
Now, we have that for every (f, g) ∈ V,
ρ(f,g)=R3kfkEk+gkAk=R3kfkÊk+fkδkjjϕE+gkÂk+gkkϕA=R3k(fkÊk+gkÂk)ϕEdiv(f)ϕAdiv(g)=R3kfkÊk+gkÂk,
as desired. The second to last line is implied by the divergence theorem, while the last line is implied by the assumption that f and g are divergence free.□

In what follows, when we have linear functionals on V determined by smooth, compactly supported field configurations (E,A)Tc1,0(R3)Tc0,1(R3)V, we will always choose E and A to be divergence free without further comment, as justified by the preceding lemma. Note that this means that we impose both the Coulomb gauge and the first of the Maxwell equations through the kinematical structure of V. The remaining Maxwell equations are encoded in the choice of complex structure.

To specify a complex structure, we define the operator μEM:Tc0,1(R3)Tc1,0(R3) by

(μEMf)jδjk(2)1/2fk

for all fTc0,1(R3). Then, we define the complex structure JEM : VV by

JEM(f,g)=(μEM1g,μEMf)

for all (f, g) ∈ V. Again, JEM is the unique complex structure compatible with inertial time translations of the fields (E, A) satisfying Maxwell’s equations. As in the Secs. V A and V B, we define the electromagnetic number operators NEM(F) as the number operators corresponding to this choice of complex structure, i.e., NEM(F)NJEM(F) for each FV.

We again consider the representation of W(V,0) as AP(V′). In this representation, we have the following form for N0EM(f,g) as a real-valued function on phase space.

Proposition 14.
For any (f, g) ∈ V,
N0EM(f,g)(E,A)=12R3kEkfk+Akgk2+12R3kAk(μEMf)kEk(μEMg)k2
for all(E,A)T1,0(R3)T0,1(R3)V.

Furthermore, we can construct the classical total electromagnetic number operatorN¯0EM by letting {Fk} be any αJEM-orthonormal basis for V and defining

N¯0EMkN0EM(Fk).

The following proposition provides an explicit form for the total number operator.

Proposition 15.
N¯0EM(E,A)=12R3kEk(μEM1E)k+Ak(μEMA)k
for any(E,A)Tc1,0(R3)Tc0,1(R3)V.

Proof.

This follows from an analogous calculation to that in the Proof of Proposition 9.□

Finally, we can construct the classical electromagnetic HamiltonianH0EM. Let {fk} be any orthonormal basis for Tc1,0(R3) with the generalized L2-inner product

f,g=R3l,mδjlfjgl.

Now, we define

H0EMkN0EM(fk,0).

With this definition, the classical limit of the electromagnetic Hamiltonian also takes a familiar form as a real-valued function on phase space.

Proposition 16.
H0EM(E,A)=12R3j,kδjkEjEk+δjkcurl(A)jcurl(A)k
for any(E,A)Tc1,0(R3)Tc0,1(R3)V.

Proof.
An analogous calculation to that in the Proof of Proposition 10 yields
H0EM(E,A)=12R3j,kδjkEjEk+δjk(2Aj)Ak,
so it suffices to show that ∫∑j,kδjk(−∇2Aj)Ak = ∫∑j,kδjkcurl(A)jcurl(A)k.
To this end, first note that since we choose A to be divergence free, we have 0 = 1A1 + 2A2 + 3A3, and hence,
A112A1=A11(2A2+3A3)A222A2=A22(1A1+3A3)A332A3=A33(1A1+2A2).
From this, consequently,
R3j,kδjk(2Aj)Ak=R3A1(12A1+22A1+32A1)A2(12A2+22A2+32A2)A3(12A3+22A3+32A3)=R3(1A2)2+(2A1)2+(2A3)2+(3A2)2+(1A3)2+(3A1)2R3A112A1+A222A2+A332A3=R3(1A2)2+(2A1)2+(2A3)2+(3A2)2+(1A3)2+(3A1)2+R3A11(2A2+3A3)+A22(1A1+3A3)+A33(1A1+2A2)=R3(1A2)2+(2A1)2+(2A3)2+(3A2)2+(1A3)2+(3A1)22R3(1A2)(2A1)+(2A3)(3A2)+(1A3)(3A1)=R3(1A22A1)2+(2A33A2)2+(1A33A1)2=R3j,kδjkcurl(A)jcurl(A)k,
where we obtain the second and fourth equalities from integration by parts.□

This shows that the classical limit of the electromagnetic Hamiltonian is the classical total energy of the electromagnetic field for initial data of compact support. This, of course, is the conserved quantity of the electromagnetic field corresponding to the inertial time-like symmetries of Minkowski spacetime.

In this paper, we have analyzed the classical limits of unbounded quantities and illustrated our methods for number operators and Hamiltonians in linear Bosonic quantum field theories. Our strategy has stayed close to the framework of strict deformation quantization by (i) looking for norm approximations and (ii) treating physical magnitudes as elements of an abstract partial *-algebra rather than focusing on particular Hilbert space representations. Using developments in the theory of algebras of unbounded operators, we considered continuous extensions of positive quantization maps. We used these extensions to prove norm approximations in the classical limit for unbounded quantities including field operators, creation and annihilation operators, and number operators. We then analyzed the classical limits of number operators and associated Hamiltonians for the Klein–Gordon field theory and the Maxwell field theory. We established that the methods developed in this paper yield a unified approach to the classical limit for both the Minkowski number operators and the Rindler number operators for the Klein–Gordon field, which are unbounded operators that are typically understood to be affiliated with unitarily inequivalent representations of the Weyl algebra. In both cases, the classical limits of the associated Hamiltonians are the classical conserved energy quantities associated with certain time-like symmetries, as expected. Similarly, we established (as expected) that the classical limit of the Hamiltonian for the free Maxwell field is the classical energy of the electromagnetic field. Thus, the methods developed here for taking classical limits of unbounded operators capture the intended use of the classical limit while extending its application beyond C*-algebras in strict quantization.

The authors thank Adam Caulton and Charles Godfrey and the audience of the conference “Foundations of Quantum Field Theory” (Rotman Institute of Philosophy, 2019) for helpful comments and discussion. B.H.F. acknowledges support from the Royalty Research Fund at the University of Washington during the completion of this work as well as the National Science Foundation under Grant No. 1846560.

1.
N. P.
Landsman
,
Foundations of Quantum Theory: From Classical Concepts to Operator Algebras
(
Springer
,
2017
).
2.
N. P.
Landsman
,
Mathematical Topics Between Classical and Quantum Mechanics
(
Springer
,
New York
,
1998b
).
3.
N. P.
Landsman
, “
Between classical and quantum
,” in
Handbook of the Philosophy of Physics
, edited by
J.
Butterfield
and
J.
Earman
(
Elsevier
,
New York
,
2007
), Vol. 1, pp.
417
553
.
4.
M. A.
Rieffel
, “
Deformation quantization of Heisenberg manifolds
,”
Commun. Math. Phys.
122
,
531
562
(
1989
).
5.
M.
Rieffel
, Deformation Quantization for Actions of Rd, Memoirs of the American Mathematical Society (
American Mathematical Society
,
Providence, RI
,
1993
).
6.
M.
Rieffel
, “
Quantization and C*-Algebras
,”
Contemp. Math.
167
,
66
97
(
1994
).
7.
S.
Waldmann
, “
States and representations in deformation quantization
,”
Rev. Math. Phys.
17
(
1
),
15
75
(
2005
).
8.
S.
Waldmann
, “
Recent developments in deformation quantization
,” in
Quantum Mathematical Physics: A Bridge Between Mathematics and Physics
(
Birkhäuser/Springer
,
Cham
,
2016
), pp.
421
439
[
Selected Papers Based on the Presentations at the International Conference, Regensburg, Germany, September 29–October 2, 2014
].
9.
S.
Hollands
and
R. M.
Wald
, “
Local Wick polynomials and time ordered products of quantum fields in curved spacetime
,”
Commun. Math. Phys.
223
,
289
326
(
2001
).
10.
S.
Hollands
and
R. M.
Wald
, “
Axiomatic quantum field theory in curved spacetime
,”
Commun. Math. Phys.
293
,
85
125
(
2010
).
11.
K.
Rejzner
,
Perturbative Algebraic Quantum Field Theory: An Introduction for Mathematicians
(
Springer
,
New York
,
2016
).
12.
M.
Fragoulopoulou
,
A.
Inoue
, and
K.-D.
Kürsten
, “
Old and new results on Allan’s GB*-algebras
,”
Banach Center Publ.
91
,
169
178
(
2010
).
13.
R.
Haag
,
Local Quantum Physics
(
Springer
,
Berlin
,
1992
).
14.
K.
Hepp
, “
The classical limit for quantum mechanical correlation functions
,”
Commun. Math. Phys.
35
,
265
277
(
1974
).
15.
Z.
Ammari
,
S.
Breteaux
, and
F.
Nier
, “
Quantum mean-field asymptotics and multiscale analysis
,”
Tunis. J. Math.
1
(
2
),
221
272
(
2019
).
16.
M.
Combescure
and
D.
Robert
,
Coherent States and Applications in Mathematical Physics
(
Springer
,
Dordrecht
,
2012
).
17.
M.
Falconi
, “
Cylindrical Wigner measures
,”
Doc. Math.
23
,
1677
1756
(
2018
).
18.
B. H.
Feintzeig
, “
The classical limit as an approximation
,”
Philos. Sci.
87
(
4
),
612
539
(
2020
).
19.
N. P.
Landsman
, “
Deformations of algebras of observables and the classical limit of quantum mechanics
,”
Rev. Math. Phys.
5
(
4
),
775
806
(
1993a
).
20.
N. P.
Landsman
, “
Strict deformation quantization of a particle in external gravitational and Yang-Mills fields
,”
J. Geom. Phys.
12
,
93
132
(
1993b
).
21.
N. P.
Landsman
, “
Twisted Lie group C*-Algebras as strict quantizations
,”
Lett. Math. Phys.
46
,
181
188
(
1998a
).
22.
N. P.
Landsman
, “
Spontaneous symmetry breaking in quantum systems: Emergence or reduction?
,”
Stud. Hist. Philos. Mod. Phys., Part B
44
,
379
394
(
2013
).
23.
E.
Binz
,
R.
Honegger
, and
A.
Rieckers
, “
Construction and uniqueness of the C*-Weyl algebra over a general pre-symplectic space
,”
J. Math. Phys.
45
(
7
),
2885
2907
(
2004a
).
24.
J.
Manuceau
,
M.
Sirugue
,
D.
Testard
, and
A.
Verbeure
, “
The smallest C*-algebra for the canonical commutation relations
,”
Commun. Math. Phys.
32
,
231
243
(
1974
).
25.
E.
Binz
,
R.
Honegger
, and
A.
Rieckers
, “
Field-theoretic Weyl quantization as a strict and continuous deformation quantization
,”
Ann. Inst. Henri Poincaré
5
,
327
346
(
2004b
).
26.
R.
Honegger
,
A.
Rieckers
, and
L.
Schlafer
, “
Field-Theoretic Weyl deformation quantization of enlarged Poisson algebras
,”
Symmetry, Integrability Geom.: Methods Appl.
4
,
047
084
(
2008
).
27.
M.
Reed
and
B.
Simon
,
Functional Analysis
(
Academic Press
,
New York
,
1980
).
28.
F.
Bagarello
,
M.
Fragoulopoulou
,
A.
Inoue
, and
C.
Trapani
, “
The completion of a C*-algebra with a locally convex topology
,”
J. Oper. Theory
56
(
2
),
357
376
(
2006
).
29.
F.
Bagarello
,
M.
Fragoulopoulou
,
A.
Inoue
, and
C.
Trapani
, “
Structure of locally convex quasi C*-algebras
,”
J. Math. Soc. Jpn.
60
(
2
),
511
549
(
2008
).
30.
F.
Bagarello
,
M.
Fragoulopoulou
,
A.
Inoue
, and
C.
Trapani
, “
Locally convex quasi C*-normed algebras
,”
J. Math. Anal. Appl.
366
,
593
606
(
2010
).
31.
M.
Fragoulopoulou
,
A.
Inoue
, and
K.-D.
Kürsten
, “
On the completion of a C*-normed algebra under a locally convex algebra topology
,”
Contemp. Math.
427
,
155
166
(
2007
).
32.
J.-P.
Antoine
,
A.
Inoue
, and
C.
Trapani
,
Partial *-Algebras and Their Operator Realization
(
Kluwer
,
Dordrecht
,
2002
).
33.
A.
Inoue
,
Tomita-Takesaki Theory in Algebras of Unbounded Operators
(
Springer-Verlag
,
Berlin
,
1998
).
34.
K.
Schmüdgen
,
Unbounded Operation Algebras and Representation Theory
(
Springer
,
Berlin
,
1990
).
35.
C. A.
Berger
and
L. A.
Coburn
, “
Toeplitz operators and quantum mechanics
,”
J. Funct. Anal.
68
,
273
399
(
1986
).
36.
S.
Waldmann
, “
Positivity in Rieffel’s strict deformation quantization
,” in
XVIth International Congress on Mathematical Physics, Prague, Czech Republic, August 3–8, 2009. With DVD
(
World Scientific
.
Hackensack, NJ
,
2010
), pp.
509
513
.
37.
R.
Honegger
and
A.
Rieckers
, “
Some continuous field quantizations, equivalent to the C*-Weyl quantization
,” in
Publications of the Research Institute for Mathematical Sciences
(
Kyoto University
,
2005
), Vol. 41, Issues 113-138.
38.
B. H.
Feintzeig
, “
On the choice of algebra for quantization
,”
Philos. Sci.
85
(
1
),
102
125
(
2018a
).
39.
B. H.
Feintzeig
, “
The classical limit of a state on the Weyl algebra
,”
J. Math. Phys.
59
,
112102
(
2018b
).
40.
K.
Sumesh
and
V. S.
Sunder
, “
On a tensor-analogue of the Schur product
,”
Positivity
20
,
621
624
(
2016
).
41.
R.
Kadison
and
J.
Ringrose
,
Fundamentals of the Theory of Operator Algebras
(
American Mathematical Society
,
Providence, RI
,
1997
).
42.
R.
Honegger
, “
On the continuous extension of states on the CCR algebra
,”
Lett. Math. Phys.
42
,
11
25
(
1997
).
43.
K.
Fredenhagen
and
K.
Rejzner
, “
Perturbative construction of models in algebraic quantum field theory
,” in
Advances in Algebraic Quantum Field Theory
(
Springer
,
2015
), pp.
31
74
.
44.
M.
Reed
and
B.
Simon
,
Methods of Modern Mathematical Physics II: Fourier Analysis, Self-Adjointness
(
Academic Press
,
New York
,
1975
).
45.
B. S.
Kay
, “
A uniqueness result in the Segal-Weinless approach to linear Bose fields
,”
J. Math. Phys.
20
,
1712
1713
(
1979
).
46.
B. S.
Kay
, “
The double-wedge algebra for quantum fields on schwarzschild and Minkowski spacetimes
,”
Commun. Math. Phys.
100
,
57
81
(
1985
).
47.
R.
Wald
,
Quantum Field Theory in Curved Spacetime and Black Hole Thermodynamics
(
University of Chicago Press
,
Chicago
,
1994
).
48.
B. S.
Kay
and
R. M.
Wald
, “
Theorems on the uniqueness and thermal properties of stationary, nonsingular, quasifree states on spacetimes with a bifurcate killing horizon
,”
Phys. Rep.
207
,
49
136
(
1991
).