Skip to main content
Springer logoLink to Springer
. 2025 Mar 25;192(4):44. doi: 10.1007/s10955-025-03415-y

Implementing Bogoliubov Transformations Beyond the Shale–Stinespring Condition

Sascha Lill 1,
PMCID: PMC11933170  PMID: 40144799

Abstract

We define infinite tensor product spaces that extend Fock space, and allow for implementing Bogoliubov transformations which violate the Shale or Shale–Stinespring condition. So an implementation on the usual Fock space would not be possible. Both the bosonic and fermionic case are covered. Conditions for implementability in an extended sense are stated and proved. From these, we derive conditions for a quadratic Hamiltonian to be diagonalizable by a Bogoliubov transformation that is implementable in the extended sense. We apply our results to Bogoliubov transformations from quadratic bosonic interactions and BCS models, where the Shale or Shale–Stinespring condition is violated, but an extended implementation nevertheless works.

Keywords: Non-perturbative renormalization, Dressing transformations, Infinite tensor product spaces, Fock space extensions, Bogoliubov transformations, Quadratic Hamiltonians

Introduction

In quantum many-body systems and non-perturbative quantum field theory (QFT), one often encounters situations in which a formal expression H for a Hamiltonian is given in the physics literature, but can a priori not be interpreted as a self-adjoint operator that generates dynamics on a Hilbert space. Over the past decades, a plethora of mathematical tools has been conceived to overcome this issue [1, 2].

An established but little investigated tool in this area are Fock space extensions, such as the infinite tensor product (ITP) framework, introduced by von Neumann [3]. There have been attempts to apply this framework to quantum electrodynamics (QED) scattering theory [46]. Rigorous results exist on the implementation of Weyl transformations (i.e., linear exponents) that are not implementable on the usual Fock space [79].

The present paper takes the step from linear to quadratic exponents, i.e., from Weyl to Bogoliubov transformations: We derive conditions which ensure that a Bogoliubov transformation, which is not implementable on the usual Fock space, can nevertheless be implemented on a suitable ITP space.

Non-implementable Weyl- and Bogoliubov transformations are both a well-researched topic in the C-algebraic formulation of quantum dynamics [10]. Therefore, they serve as an ideal test area for mathematical tools that may ultimately turn out advantageous for more sophisticated operator transformations.

Let us explain a bit more precisely, how the implementation of Bogoliubov transformations via ITPs works. Roughly speaking, a Bogoliubov transformation V=uvv¯u¯ replaces annihilation operators a(f),a(f) by

b(f)=a(uf)+a(vf¯),b(f)=a(vf¯)+a(uf),

with f being an element of the one-particle Hilbert space h, where uv are linear operators on h and where f¯ is the complex conjugate of f. This replacement may diagonalize quadratic Hamltonians H [1114]. Related transformations allow for eliminating inconvenient terms of higher order in more sophisticated H [1518].

It is desirable to find a unitary operator UV on Fock space F, such that UV establishes the replacement ab via

UVa(f)UV=b(f),UVa(f)UV=b(f). 1

In that case we say that UV implements the transformation V and we call V “implementable” (in the regular sense). It is well-known [19, 20] that V is implementable, if and only if the Shale condition (bosonic case) or the Shale–Stinespring condition (fermionic case) holds, which asserts that tr(vv)<. Situations with non-implementable Bogoliubov transformations occur, for instance, in relativistic models [21, 22], and within many-body systems of infinite size [2325].

We prove that under certain conditions, V is nevertheless implementable on certain ITP spaces H^=kNHk, which extend all of F, while being a sum of typically uncountably many spaces that are naturally isomorphic to F. We further elucidate the structure of H^ in the end of Sect. 2.2, as well as in Appendix A.

Also, the way how we diagonalize formal Hamiltonians H by implementers UV differs from the usual procedure on Fock space: For some formal H consisting of a product of a- and a-operators, we aim at defining (see Fig. 1)

H~=UV-1(H+c)UV 2

on a dense subspace DFF, where H~ is the version of H with a replaced by b and with normal ordering applied. The “renormalization constant” c stems from normal ordering and can be infinite. While (2) may not always be achieved, we can still split H=nNH(n) and c=nNc(n), such that

H~=nNUV-1H(n)+c(n)UV. 3

That is, we define an operator UV:DFH^ such that (H(n)+c(n)) maps the space UV[DF]H^ into itself. Further, H~ allows for a self-adjoint extension and can thus be seen as the well-defined renormalized version of H.

Fig. 1.

Fig. 1

Renormalization of H using ITPs

Our main result is that in the extended sense, specified in Definition 5.1, V can indeed be implemented on H^ in the bosonic (Theorem 5.5) and the fermionic case (Theorem 5.6) if the spectrum of the operator vv is countable. We may then naturally extend the implementer to a unitary operator UV:H^H^, see Remark 5.7.

It is worth mentioning that within the extended state space (ESS) framework [26], a similar result was recently proven even for arbitrary vv in the bosonic case [27]. In Appendix B, we briefly discuss a much simpler ESS construction for vv having discrete spectrum, which allows for an extended implementation in both the bosonic and fermionic case (Propositions B.3 and B.4).

With respect to ITPs, the fermionic ESS implementation requires the additional restriction to a finite number of particle–hole transformed modes. Nevertheless, we expect a fermionic ESS construction, similar to [27], to also be successful for generic vv.

By contrast, the ITP construction cannot be expected to achieve an implementation for generic vv: Still, the ITP space would be well-defined as xXHx with X being a possibly uncountable set related to σ(vv). However, a(f)=xf(x)ax would only be defined for countable sums in x. So a(f) would be ill-defined unless f(x) is everywhere 0 apart from countably many x.

The ultimate goal would be to implement more general operator transformations W such that

H~=W-1(H+c)W, 4

is well-defined on DFF and allows for a self-adjoint extension. Here, c is a general counterterm and not necessarily just a constant. Transformations W as above arise from non-perturbative cutoff renormalization [2833], when formally removing the IR- or UV-cutoff from the employed dressing transformations. In contrast to cutoff renormalization, the direct renormalization by Fock space extensions does not involve limiting processes or cutoffs that break Lorentz invariance. A similar cutoff-free non-perturbative renormalization technique, also known as “interior–boundary conditions” (IBC), has recently been proposed and investigated [3442]. However, IBC renormalization is limited to cases where the free and interacting Hamiltonian can be defined in the same CCR/CAR representation. This can be a severe restriction, for instance, in relativistic models,1 and renormalization by Fock space extensions is designed to overcome this issue.

The rest of this paper is structured as follows: In Sect. 2, we give the basic definitions of second quantization and the ITP framework. Section 3 recaps known properties on Bogoliubov transformations that are needed for the extended implementation. In Sect. 4, we define the V-dependent ITP spaces H^ and prove that creation and annihilation operators are well-defined on them (Lemma 4.9). On these ITP spaces, we define implementability in Sect. 5, and prove that it is satisfied (Theorems 5.5 and 5.6). Section 6, is devoted to the diagonalization of Hamiltonians in the extended sense (Propositions 6.3 and 6.4). In Sect. 7 we examine two examples for a diagonalization in the extended sense.

Basic Definitions

Fock Space Notions

We consider a measure space (X,μ) with XRd, where we focus on X=Rd and X=N. A configuration of NN0 particles is given by the tuple q=(x1,,xN), which is an element of the ordered configuration space

Q(X):=N=0Q(X)(N):=N=0XN. 5

Q(X) allows for a standard topology and a measure μN on each sector Q(X)(N), hence yielding a topology and a measure μQ on Q(X). The full Fock space is then

F(X):=L2(Q(X),μQ). 6

The corresponding scalar product is Φ,Ψ:=Q(X)Φ(q)¯Ψ(q)dq with Φ,ΨF(X) and the overline denoting complex conjugation. For ΨC0(Q(X)), a unique continuous representative function exists. This includes smooth functions ΨC(Q(X)) and smooth functions with compact support ΨCc(Q(X)).

For describing bosonic/fermionic particle exchange symmetries, we introduce the symmetrization operators S+,S-:F(X)F(X) defined by

(S±Ψ)(x1,,xN):=1N!σSN(±1)(1-sgn(σ))/2Ψ(xσ(1),,xσ(N)), 7

with permutation group SN. Here, (1-sgn(σ))/2 is 0 if the permutation is even, and 1 if it is odd. The bosonic (+) and fermionic (-) Fock space is given by

F±(X):=S±[F(X)]. 8

The (N) -sectors of Fock space are F(X)(N):=L2(Q(X)(N),C), with symmetrized and antisymmetrized analogues F±(N). Note that we may equivalently write

F(X):=N=0F(X)(N),F(X)(N):=hhNtimes, 9

with h:=L2(X). A change of basis then allows us to identify h with 2=L2(N) (or a subspace thereof), thus identifying F(X) with F(N). In the following, we drop the (X) if not explicitly needed.

Creation and annihilation operators a(f),a(f) for some fh can be defined by using q\xjXN-1 for denoting the removal of one particle xj from a configuration qXN:

(a±(f)Ψ)(q)=j=1N(±1)jNf(xj)Ψ(q\xj),(a±(f)Ψ)(q)=N+1f(x)¯Ψ(q,x)dμ(x). 10

It is well-known that the fermionic operators a-,a- are bounded and hence defined on all ΨF-, while the bosonic a+,a+ are unbounded, but can still be defined on a dense subspace of F+. Further, (10) implies the canonical commutation/anticommutation relations (CCR/CAR):

a±(f),a±(g)±=f,gh,a±(f),a±(g)±=0=a±(f),a±(g)±, 11

with commutator [A,B]+=[A,B]=AB-BA and anticommutator [A,B]-={A,B}=AB+BA. In the following, we will drop the indices ± if there is no risk of confusion.

It is also customary to just consider a(f),a(f) not as operators, but as formal expressions within a * -algebra2

A=A±generatedbya±(f),a±(f)fh. 12

The involution is given by a(f)=a(f) and the multiplication in A is such that the CCR/CAR hold. In particular, A- is a C-algebra by boundedness of operators.

Infinite Tensor Products

In this subsection, we give a quick introduction to general ITPs, as introduced by von Neumann [3], see also [44]. Some useful lemmas and remarks concerning this construction are given in Appendix A. For a more thorough discussion, we refer the reader to [3].

We consider a (possibly uncountable) index set I, and for each kI a Hilbert space Hk with scalar product ·,·k and induced norm ·k. The aim is to construct a vector space, which is generated formally by ITPs

Ψ=kIΨk, 13

or equivalently, by families (Ψ)=(Ψk)kI,ΨkHk. If I is countable, then (Ψ)=(Ψ1,Ψ2,) defines a sequence. Each family gets assigned the formal expression

(Ψ):=kIΨkk, 14

which we will later use for defining a norm. In order to answer the question, whether (14) defines a complex number, one introduces the notions of convergence within a (possibly uncountable) sum:

  • For zkC,kI, we call kIzk or kIzk convergent to aC, if for all δ>0, there exists some finite set Iδ, such that for all finite sets JI with IδJ, we have
    a-kJzkδora-kJzkδ,respectively. 15

A simple consequence of this definition is that kIzk can only converge if zk0 occurs for only countably many kI. So the question of convergence reduces to that of sequence convergence. Further, it is shown in [3] that kIzk< if and only if we have zk=0 for at least one kI or if kI|zk-1|<. The heuristic reason is that kIzk=expkIlnzk and lnzk can be linearly approximated near 1 as lnzk=1-zk+O((1-zk)2).

In case kI|zk| converges to a nonzero number, then kIzk converges if and only if no infinite phase variation occurs. That is, if arg(zk)(-π,π] is the phase of the complex number zk, then it is required that

kI|arg(zk)|<. 16

In order to establish a notion of convergence, even when (16) is violated, one defines that kIzk is quasi-convergent if and only if kI|zk| converges. A family (Ψ)=(Ψk)kI is now called a

  • C-sequence ((Ψ)Cseq) if and only if kIΨkk<,

  • C0-sequence if and only if kIΨkk-1<kIΨkk2-1<.

Each C0-sequence is also a C-sequence. For all C-sequences, we have a well-defined value (Ψ)C by (14) and each C-sequence, that is not a C0-sequence, must automatically satisfy (Ψ)=0. However, (14) only defines a seminorm, since there exist (Ψ)0 with kIΨkk=0. To make it a norm, one defines

  • Inline graphic as the space of all functionals on Cseq which are conjugate-linear in each component.

Following [3], we can embed Inline graphic by identifying (Φ)Cseq with the functional

Φ=ι((Φ)):(Ψ)kIΦk,Ψkk. 17

This identification essentially sets up an equivalence relation C on Cseq, where (Φ)C(Φ), whenever ι((Φ))=ι((Φ)). In Proposition A.1, we show that equivalence is given if and only if (Φ) and (Φ) just differ by a family of complex factors (ck)kI with kIck=1. The functionals in ι[Cseq] are then the equivalence classes and the span of these functionals is denoted by [3]:

  • ~kIHk:=span(ι[Cseq]).

In the following, we may drop the embedding map ι and simply identify (Φ) with Φ, whenever the identification is obvious. An inner product ·,· can uniquely be defined on ~kIHk via

Φ,Ψ=kIΦk,Ψkk, 18

which is a quasi-convergent product that we understand to be 0 whenever it is not convergent. ·,· makes ~kIHk a pre-Hilbert space and induces a norm Φ agreeing with (14) under identification Φ=(Φ). This norm allows completing ~kIHk to a Hilbert space:

  • The infinite tensor product space H^=kIHk is defined as the space of all Inline graphic, such that there exists a Cauchy sequence (Φ(r))rN~kIHk with respect to · that converges to Φ in the weak-* topology on Inline graphic.

One may indeed extend ·,· to H^, making the latter a Hilbert space [3]. In order to further analyze its structure, we divide H^ into subspaces, For which we split the set of C0-sequences into equivalence classes via

  • equivalence: (Φ)(Ψ):kIΦk,Ψk-1<

  • weak equivalence: (Φ)w(Ψ):kI||Φk,Ψk|-1|<

The respective equivalence classes are called C and Cw, and the corresponding linear spaces of an equivalence class are

  • kICHk:=span{Ψ(Ψ)C:ι((Ψ))=Ψ}¯· for equivalence

  • kICwHk:=span{Ψ(Ψ)Cw:ι((Ψ))=Ψ}¯· for weak equivalence

Now, each C0-sequence (Ψ) in some equivalence class [(Ω)]=C (with (Ω)Cseq being interpreted as the vacuum vector) can be written in coordinates as follows [3, Theorem V]: Choose an orthonormal basis (ek,n)nN0 for each Hk, such that Ωk=ek,0 (we think of ek,0 as mode k being in the vacuum). Then, (Ψ)=(Ψk)kI is uniquely specified by the coordinates ck,n:=ek,n,ΨkkC. In this coordinate representation, it is true that

  • kICHk is the closure of the space spanned by all normalized C0-sequences, where ck,0=1 for all but finitely many kI.

Or heuristically speaking, “almost all Ψk are in the vacuum”. By [3, Theorem V], also a generic ΨkICHk can be written as

Ψ=n(·)Fa(n(·))kIek,n(k), 19

with F being the countable set of all functions n:IN0 with n(k)=0 for almost all kI, and a(n(·))C being the coordinates of Ψ with n(·)F|a(n(·))|2<.

Bogoliubov Transformations

In this section, we introduce our notation for Bogoliubov transformations and recap some important properties. Standard references on the subject are [45] and [46].

Transformation on Operators

Consider the one-operator subspace W1 of A, which is linearly spanned by a±(f),a±(g), f,gh. By an algebraic Bogoliubov transformation, we mean any bijective map VA:W1W1, which sends a±(f),a±(g) to a new set of creation and annihilation operators b±(f),b±(g), such that b±(f) is the adjoint of b±(f) and the CAR/CCR are conserved under VA and its adjoint.

To facilitate the presentation, we fix a basis (ej)jNh which identifies fh with an equally denoted vector f=(fj)jN2 by fj:=ej,f. This way, we can also identify a(f)+a(g¯)W1 with a vector (f,g)22, so an algebraic Bogoliubov transformation VA is identified with a linear operator V on 22, which we just call “Bogoliubov transformation”. We may also encode sums of creation and annihilation operators by vector pairs (f,g)22 via the generalized creation/annihilation operators

A±:22A±,(f1,f2)a±(f1)+a±(f2¯)=jf1,ja±(ej)+f2,ja±(ej),A±:22A±,(g1,g2)a±(g1)+a±(g2¯)=jg1,j¯a±(ej)+g2,j¯a±(ej). 20

A Bogoliubov transformation is then encoded by a 2×2 block matrix

V=uvv¯u¯, 21

with operators u,v:22. The case of unbounded uv is treated later in Sect. 4. The Bogoliubov transformed operators are then given by

b±(f)=A±(V(f,0))=a±(uf)+a±(vf¯),b±(g)=A±(V(g,0))=a±(vg¯)+a±(ug). 22

In order for V to be a Bogoliubov transformation, we require that both V and V conserve the CAR/CCR, so

b±(f),b±(g)±=f,g,b±(f),b±(g)±=0=b±(f),b±(g)±, 23

and the same, if in (22) V is replaced by V. An explicit calculation shows that this conservation is equivalent to the Bogoliubov relations

uuvTv¯=1,uvvTu¯=0,uuvv=1,uvTvuT=0, 24

with (uT)ij=uji,(u¯)ij=uij¯,(u)ij=uji¯ and the same for vij. Since vv is self-adjoint, we have vv=vv¯=vTv¯, so the first equation is equivalent to uuvv=1. The generalized creation and annihilation operators also allow for particularly easy “generalized CAR/CCR”: Using the standard scalar product on F,G22:

F,G=f1f2,g1g2=jf1,j¯g1,j+f2,j¯g2,j, 25

and S-=id, S+=100-1, we obtain the generalized CAR/CCR:

A±(F),A±(G)±=F,S±G,A±(F),A±(G)±=A±(F),A±(G)±=0. 26

Implementation on Fock Space

In the above notation, a Bogoliubov transformation is implementable (in the regular sense), if there exists a unitary operator UV:FF, such that

UVA(F)UV=A(VF). 27

We recap some of the basic steps of the implementation process presented in [45] for tr(vv)<, as they have to be carried out in a slightly modified way for an implementation in the extended sense.

The main task within the implementation is to find the Bogoliubov vacuum ΩVF, which is the vector annihilated by all operators b(f):

b(f)ΩV=a(uf)+a(vf¯)ΩV=0f2. 28

If we can find such an ΩV, then it is an easy task to transform any product state vector a(f1)a(fn)ΩF via:

UVa(f1)a(fn)Ω=b(f1)b(fn)ΩV. 29

The span of these state vectors (also called algebraic tensor product) is dense within F, so we can transform any ΨF by means of (29).

To find ΩV, we decompose the transformation V into modes by constructing vectors fj, such that ufj,vfj¯ are proportional to the same normalized vector gj2, i.e.

Vfj0=ufjv¯fj=μjgjνjgj¯UVa(fj)UV=μja(gj)+νja(gj), 30

where μj,νjC. If (30) holds, we only have to solve (νja(gj)+μja(gj))ΩV for each gj, separately.

Bosonic Case

Here, (30) can indeed be fulfilled: Following [45], introducing the complex conjugation Jh~j=Jh~j=h~j¯, the operator C:=uvJ has an eigenbasis (fj) with Cfj=λjfj,λjC. One may further show that gj:=μj-1ufj defines an orthonormal basis and that vfj¯=vJff=νjgj where μj,νj0 are fixed by

μj2-νj2=1,λj=μjνj. 31

The condition (μja(gj)+νja(gj))ΩV=0 leads to one recursion relation per mode, which is formally solved by

ΩV=j1-νj2μj21/4exp-jνj2μj(a(gj))2Ω. 32

The Shale condition now indicates when ΩV lies in Fock space, which is if and only if jνj2μj2<jνj2<. In that case, the transformation is implemented by [15, (3.1)]:

UV=exp-jξj2((a(gj))2-(a(gj))2)Ugf=:jNUj,V, 33
withsinhξj:=νjcoshξj:=μj, 34

and where Ugf:FF is the unitary basis change transformation

Ugf:fj1fjngj1gjnj1,,jnN. 35

For a more general discussion on UV, we refer the reader to [10, Theorem 16.47].

Fermionic Case

In the fermionic case, following [45], we can find a common orthonormal eigenbasis (fj)jJ of CC (eigenvalues λj2), of uu (eigenvalues μj2), and of vv (eigenvalues νj2=1-μj2). The index set can be split as J=JJN with J containing all indices of zero eigenvectors and J those of nonzero eigenvectors. Further, the jJ can be rearranged in pairs as J:={jj=2ij=2i-1,iI} such that

Cf2i=λ2if2i-1,Cf2i-1=-λ2if2i, 36

for some IN. For jJ, two cases may occur: If νj=1, then we have a particle–hole transformation, for which we write jJ1 and get

Vfj0=0ηj¯,jJ1, 37

for a suitable choice of the phase ηj=eiφvfj¯. The case νj=0 will be denoted by jJ0=J\J1, and we have

Vfj0=ηj0,jJ0, 38

for a suitable phase choice of ηj=eiφufj. In case jJ with iIλ2i0,μ2i0 (Cooper pair), we may define the normalized vectors

η2i:=αi-1uf2i,η2i-1:=αi-1uf2i-1, 39

where αi,βi>0, αi=μ2i=μ2i-1, αi2+βi2=1, and for which

Vf2i0=αiη2iβiη2i-1¯,Vf2i-10=αiη2i-1-βiη2i¯,iI. 40

Relations (37), (38) and (40) now replace (30). The Bogoliubov vacuum is

ΩV=jJ1a(ηj)iIαi-βia(η2i)a(η2i-1)Ω, 41

and the implementer is

UV=jJ1a(ηj)+a(ηj)exp-iIξia(η2i)a(η2i-1)-a(η2i-1)a(η2i)UηfUV=:jJUj,ViIU2i,2i-1,V,withsinξi:=βicosξi:=αi, 42

and where Uηf is the unitary basis change transformation

Uηf:fj1fjnηj1ηjnj1,,jnJ. 43

Bogoliubov Transformations: Extended

In the extended case, v is possibly unbounded, so we must show that the Bogoliubov relations (24) survive the extension. We do this in Sect. 4.1, while preparing the spectral decomposition for the final ITP construction. In Sect. 4.2, we define an extended * -algebra A¯e of creation and annihilation operator products, and in Sect. 4.3, we finish the construction of H^, with respect to a given Bogoliubov transformation V.

Extension of the Bogoliubov Relations

Throughout the following construction, we will assume that vv is densely defined and self-adjoint. In that case, we can define the self-adjoint operators CC:=vv(1±vv) and |C|=CC by spectral calculus. By the spectral theorem in the form of [47, Thm. 10.9], we may then decompose 2 as a direct integral

2=σ(|C|)Cndμ1(λ), 44

where σ(|C|)=σ is the spectrum of |C| , μ1 is a suitable measure on it and n:σN{} [47, Def. 7.18] is a measurable dimension function. Put differently, as visualized in Fig. 2, we can find a spectral set X=λσ{λ}×YλR2 with YλZ,|Yλ|=n(λ) accounting for multiplicity and unitary maps3

UXf:2L2(X),UfX=UXf-1, 45

such that

|C|=UfXλUXf, 46

with λ being the operator on L2(X) that multiplies by λ(x). In addition, we denote Y=λσYλZ with |Y| being an upper bound for the multiplicity of any eigenvalue. Note that the λ here correspond to the λj in Sect. 3.2.

Fig. 2.

Fig. 2

The spectral set X for a generic spectrum of |C|

We also make use of the formulation [47, Thm. 10.4] of the spectral theorem, which provides us with a projection-valued measure P|C|, such that

|C|=Xλ(x)dP|C|(x)=σ×YλdP|C|(λ,y). 47

Here, we choose YZ, so XR2 consists of “lines” with distance 1.

The case λ=0 will turn out to be critical, whence we define

Xcrit:={x=(λ,y)Xλ=0},Xreg:=X\Xcrit. 48

Our (dense) space of test functions on the spectral set is given by:

DX:=Cc(Xcrit)Cc(Xreg). 49

The corresponding test function space in 2 is

D|C|:=UfXDX. 50

For non-open X, we interpret (49) in the same way as the definition of E(X) (131):

Cc(X):=Cc(R2)/{ϕϕ(x)=0xX}.

Lemma 4.1

(Bogoliubov relations (24) survive the extension) Suppose, u and v are defined on a common dense domain D2, such that vv is densely defined and self-adjoint, and such that the linear operator

V=uvv¯u¯,V:DD22, 51

defines a Bogoliubov transformation. That means, both V and V=uvTvuT preserve the CAR/CCR, see (20), (23).

Then u,v,u¯,v¯,u,v,uT and vT are well-defined on all of D|C|, constructed above (50). Further, the Bogoliubov relations (24) hold as a weak operator identity on D|C|. Conversely, the extended Bogoliubov relations imply conservation of the CAR/CCR under both V and V.

Proof

Well-definedness of uv on D|C| follows from the polar decompositions

v=Uv|v|,u=Uu|u|, 52

with unitary Uv,Uu:L2(X)2. The operators |v|=vv and |u|=uu=1±vv on L2(X) are bounded in the fermionic case (-), so u and v are defined on all of 2. In the bosonic case (+), they are spectral multiplications by

ν(λ)=-12+14+λ2andμ(λ)=12+14+λ2, 53

which are bounded on each bounded interval in λ. Therefore, |v| and |u| map DX into itself, and by definition (50) of D|C|, the operators v and u map D|C|2. Well-definedness of u¯ and v¯ on D|C| follows analogously and the domains of the adjoints u,v,uT and vT contain the domain of the respective original operators, so they all contain D|C|. The CAR/CCR conservation then follows by a direct computation. Checking that (24) indeed holds as weak operator identities on D|C| is straightforward to check.

Lemma 4.2

Let vv be self-adjoint. Then C=uvJ and CC are well-defined operators on D|C|, where J denotes complex conjugation. The spectrum of C is contained in the real axis for bosons and the imaginary axis for fermions.

Further, if vv has countable spectrum, then also C,CC and |C| have countable spectrum, see Fig. 3.

Fig. 3.

Fig. 3

Left: Discrete spectrum of |C| in the bosonic case. Right: In the fermionic case, |C|12 holds

Proof

By the Bogoliubov relations, CC=vv±(vv)2 holds, wherever it is defined. Now, λλ±λ2 is smooth apart from the critical points 0 (bosons) or 0 and 1 (fermions). The condition ϕD|C| means that the corresponding spectral function ϕX has compact support and is smooth apart from the critical points. This property is preserved by an application of CC, so CC:D|C|D|C| is well-defined. Hence, also |C|=CC is well-defined, and by a polar decomposition also C=UC|C| with UC:22 unitary.

In the bosonic case, C is symmetric by the Bogoliubov relations, so CC=C2 and further, σ(C) is a subset of the preimage of σ(C2)[0,) under the complex map zz2. This preimage is contained within the real axis.

In the fermionic case, the Bogoliubov relations imply C=-C, so CC=-C2. That means, σ(C) lies within the preimage of σ(C2)[0,) under the map z-z2, which is contained within the imaginary axis.

Now, suppose σ(vv) is countable. Then, σ(|C|) is the image of σ(vv) under the map zz(1±z), which sends at most 2 arguments to the same value, so σ(|C|) is also countable.

Remark 4.3

For fermions, uu+vv=1 implies that vv is bounded, so vv and also uu can be defined on all of 2. Hence, also v and u are defined on all of 2, so the Bogoliubov relations (24) hold as a strong operator identity on 2.

Remark 4.4

For bosons, it is not obvious that the Bogoliubov relations (24) hold as a strong operator identity on a dense domain of 2. In fact, it does not always hold as a strong operator identity on D|C|: As a counter-example, let vej:=jej, with respect to the canonical basis (ej)jN and let u=Uu|u| such that Uue1=cjj-1/2-εej for some ε>0. A direct calculation then shows

uve1=|u|Uue1=j1+j2j-1/2-εej. 54

For ε1, this is obviously not in 2, so uv is ill-defined on e1D|C|.

Extension of the Operator Algebra

In Sect. 2.1, we defined a * -algebra (bosonic) or C-algebra (fermionic) A generated by a±(f),a±(f),fh. On our ITP spaces, we will encounter formal expressions in creation and annihilation operators that belong to a larger algebra A¯e, which is defined with respect to a basis e=(ej)jN2. We introduce the shorthand notations aj:=a(ej),aj:=a(ej), and consider the set of finite operator products

Πe:=aj1ajmjN. 55

Then A¯e is defined as the set of all complex-valued maps

A¯e:={H:ΠeC}. 56

We formally write its elements as infinite sums

H=mNj1,,jmNHj1,,jma1am. 57

A¯e is made a * -algebra by the involution

graphic file with name 10955_2025_3415_Equ58_HTML.gif 58

It is easy to see that A¯e extends A. Resolving each a(fj) with respect to the basis e, we obtain a countable sum of the form (57), that contains each term a1am at most once.

Of particular interest will be elements of A¯e corresponding to finite sums. For a(ϕ)=jNϕjaj,ϕ2, this sum is finite if and only if

ϕDe:=ϕ2ϕj=0forallbutfinitelymanyjN. 59

Here, the index e in De emphasizes that we are working with respect to the basis (ej)jN. If (ej)jN is an orthonormal eigenbasis of |C| , then De=D|C|, since both domains are spanned by finite linear combinations of eigenvectors of CC.

Final ITP Construction and Operator Lift

Next, we fix the final definitions for our ITP spaces H^ and define products of a(ϕ),a(ϕ) with ϕDe on suitable subspaces of them. Within these definitions, we assume that vv has countable spectrum, so Lemma 4.2 applies and |C| has countable spectrum.

As argued around (47), Xσ×Y is countable and there exists an orthonormal eigenbasis (fj)jN.

For bosons we follow the construction of [45], replacing the argument “C is Hilbert–Schmidt” by “ |C| has countable spectrum”, which renders an orthonormal basis g=(gj)jN as in Sect. 3. The construction uses that by Lemma 4.1, the Bogoliubov relations still hold as a weak operator identity. Now, g takes the role of e in De and A¯e.

Definition 4.5

The bosonic infinite tensor product space is given by

H^=kNHk=kNF({gk}), 60

Note that the sequence (ek,n)nN0 of n-particle basis vectors is a canonical basis of each one-mode Fock space Hk, and can be used to describe elements of H^.

For fermions the construction of [45], with ”CC is trace class” replaced by “ |C| has countable spectrum” (which is true by Lemma 4.2), yields an orthonormal basis (ηj)jJ with countable JN as in Sect. 3.2.2. Here, η takes the role of e in De and A¯e.

The ITP construction is then a bit more delicate, since for each Cooper pair iI (so jJ), we must introduce a separate Fock space F({η2i-1})F({η2i})C4 (see Remark 5.8). We index all jJ and iI by a corresponding k(i) or k(j) , such that all kN are used and take the tensor product over those k:

Definition 4.6

The fermionic infinite tensor product space is given by

H^=kNHk=jJF({ηj})iIF({η2i-1})F({η2i}). 61

For a one-mode Fock space Hk(j):=F({ηj}), the pair (ek,0,ek,1) forms a basis of each Hk, while for a two-mode Fock space Hk(i):=F({η2i-1})F({η2i}), such a basis is given by (ek,0,0,ek,1,0,ek,0,1,ek,1,1), with the two indices denoting the particle numbers per mode.

Our next challenge is to lift the one-mode creation and annihilation operators aj,aj defined on the one- or two-mode Fock space Hk to H^.

Lemma 4.7

Consider a (possibly unbounded) operator Aj,j:Hjdom(Aj,j)Hj. Then, for Ψj(m)dom(Aj,j),

AjΨ(m):=Ψ1(m)Ψj-1(m)Aj,jΨj(m)Ψj+1(m), 62

is independent of the choice of a C-sequence (Ψ(m))=(Ψk(m))kN representing Ψ(m), and defines an operator Aj by linearity on

Ψdom(Aj):={Ψ=mMdmΨ(m)H^|mMdmAjΨ(m)<}, 63

where MN, dmC and Ψ(m) being such that AjΨ(m) is well-defined by (62).

Proof

For a fixed choice of (Ψ(m)) representing Ψ(m) such that Ψj(m)dom(Aj,j), well-definedness of AjΨ(m) is easy to see. By Lemma A.2, we can now represent Ψ=mMdmΨ(m). And if mMdmAjΨ(m) converges, then it is independent of the representation, since Aj is linear. So dom(Aj) and AjΨ are well-defined.

It remains to prove that AkΨ(m) (and hence AkΨ) is independent of the choice of (Ψ(m)) representing Ψ(m). For mM, consider a second representative C-sequence (Ψ~(m)) with Ψ~(m)=Ψ(m). By Proposition A.1, Ψ~k(m)=ckΨk(m) for some ckC with kNck=1. By linearity, Aj,jΨ~k(m)=ckAj,jΨk(m), so also AjΨ(m) and AjΨ~(m) defined by (62) just differ by the sequence of complex factors (ck)kN with kNck=1. Hence, according to Proposition A.1, they correspond to the same functional AjΨ~(m)=AjΨ(m).

The operators aj,aj may be unbounded. We thus need to carefully choose a domain to make them bounded, by restricting the space of allowed Ψ.

Definition 4.8

In the bosonic case, the space S with rapid decay in the particle number is defined as

S:=ΨnNkN0dom(Nkn)H^|NknΨck,nΨkN,nN0, 64

where ck,n>0 are suitable constants for each n and k, and Nk is the number operator on Hk, lifted to H^. The lift is possible by Lemma 4.7, which also allows defining Nkn.

In the fermionic case, the maximum particle number per mode is 1, so we always have rapid decay and simply set

S:=H^. 65

Lemma 4.9

(Products of a,a are well-defined on the ITP space) Consider the ITP space H^S corresponding to the basis (ej)jN, which is (gj)jN (bosonic) or (ηj)jN (fermionic). Then we can lift aj and aj to H^ and for ϕDg or Dη (defined in (59)), the expressions

a(ϕ)=jϕjaj,a(ϕ)=jϕj¯aj, 66

define linear operators a(ϕ):SS and a(ϕ):SS.

Proof

First, note that the sum over j in (66) is finite by definition of Dg and Dη.

In the fermionic case, aj,aj are bounded. So by [3, Lemma 5.1.1], we can lift them to bounded operators on S=H^, and also the finite linear combination in (66) is a bounded operator on S.

In the bosonic case, where j=k, we have

akΨ=Nk+1Ψ(Nk+1)Ψ(ck,1+1)Ψ, 67

so

a(ϕ)Ψk:ϕk0|ϕk|(ck,1+1)Ψ=:c1Ψ, 68

where the sum over k contains only finitely many nonzero terms, so we may call it c1>0. Hence, a(ϕ)ΨH^ is well-defined. It remains to establish rapid decay. Now,

NknakΨ=NknNk+1ΨNkn+1Ψ+NknΨ(ck,n+1+ck,n)Ψ(ck,n+1+ck,n)akΨ, 69

so by summing over k, the rapid decay condition is again satisfied and a(ϕ)ΨS.

For a(ϕ)ΨS, the same finite-sum argument can be used to obtain a(ϕ)ΨH^. However, verifying rapid decay needs a bit more attention, since akΨΨ in (69) does not generalize to ak, see also Remark 4.10. However, for all k with akΨ0 and ϕk0, there is a fixed ratio ΨakΨ=:dk>0. So with d:=maxkdk,

NknakΨ(ck,n+1+ck,n)Ψd·(ck,n+1+ck,n)akΨ. 70

For akΨ=0, the inequality is trivial. A finite sum over k then establishes a(ϕ)ΨS.

Remark 4.10

The condition ϕDf is indeed necessary, meaning we may not just allow any ϕ2 inside a(ϕ), as the following counterexample shows: For the bosonic case (j=k), consider ϕk=1k, so ϕ2\Df. For each mode k, consider the coherent state Ψk defined sector-wise by

Ψk(Nk)=e-αk2(Nk!)-12αkNk2, 71

where all αkR are set equal to the same αk=α>0 and where Ψkk=1. Then, define the ITP Ψ=kNΨk. It is easy to see that Ψ satisfies the rapid decay condition (64), as for each Ψk, Ψk(Nk)k decays exponentially in Nk. But still, (αk)kN2, so we may think of Ψ as a “coherent state with a large displacement”, living outside the Fock space. It is a well-known fact about coherent states that akΨ=αΨ, so

a(ϕ)Ψ=kϕk¯αΨ=αk1kΨ=. 72

Hence, a(ϕ) is ill-defined on Ψ.

The same happens with any coherent state product (71) and any ϕ, where kϕk¯αk=. In particular, the space of allowed (ϕk)kN is dual to the one of allowed (αk)kN.

Remark 4.11

It is also possible to define a(ϕ) for more general ϕ, if one restricts the space S. For instance, it is not too difficult to see that imposing uniform rapid decay via

ΨSuni:=ΨH^|NknΨcnΨn,kN 73

yields a well-defined product a(ϕ1)a(ϕn)ΨH^ for any ϕ1,,ϕn1.

Alternatively, one may allow for ϕ1,,ϕnp, p(1,2] by restricting to

ΨSq:={ΨH^|(Nk+1)n/2ΨcknΨwithkckq<nN}, 74

where q is the Hölder dual of p, i.e., 1p+1q=1. Note that SuniSqS.

Remark 4.12

It is possible to view the subspace kNCHk of the equivalence class C as the original Fock space with respect to the vacuum Ω=kNek,0: Recall that each ΨkNCHk can be written in coordinates as (19):

Ψ=n(·)Fa(n(·))kNek,n(k), 75

with F containing all sequences (n(k))kN, such that n(k)=0 for almost all k. Hence, each kNek,n(k) is a tensor product state of finitely many particles. Since the Fock norm and the H^-norm coincide, the vector kNek,n(k) can be seen as a Fock space vector normalized to 1. The linear combination (75) with n(·)|a(n(·))|2 can hence also be interpreted a Fock space vector.

Conversely, each Fock space vector can be written as a countable sequence (75), since the span of the above-mentioned tensor product states is dense in F.

Implementation: Extended

We proceed with defining implementability of V by an extended operator UV on H^. Lemma 5.3 establishes that UV is well-defined and Lemma 5.4 gives conditions for when UV is an implementer in the extended sense. We then prove our main results by verifying these conditions in Theorem 5.5 for bosons and Theorem 5.6 for fermions.

Definition of Extended Implementation

The implementer UV is defined on a dense subspace of Fock space DFF, that contains a finite number of particles from the space Df (defined by (59) with e=f):

DF:=spana(ϕ1)a(ϕN)Ω,NN0,ϕDf. 76

The operator UV now maps from DF into an ITP space H^. Note that due to Lemma 4.9, the expressions b(ϕ):=a(uϕ)+a(vϕ¯) and b(ϕ):=a(uϕ)+a(vϕ¯) define operators SS, i.e., in an extended sense.

Definition 5.1

We say that a linear operator UV:DFH^ implements a Bogoliubov transformation V in the extended sense, if for all ϕDf,ΨUV[DF], we have that

UVa(ϕ)UV-1Ψ=b(ϕ)Ψ,UVa(ϕ)UV-1Ψ=b(ϕ)Ψ. 77

This requires, of course, that UV-1 is well-defined. So before establishing (77), we have to show that UV is invertible. This will be one main difficulty within the upcoming proofs.

The implementer UV is constructed as follows: First we define some new vacuum vector ΩV=UVΩS, such that

b(ϕ)ΩV=0. 78

Then we make UV change a- into b-operators:

Definition 5.2

Given a Bogoliubov transformed vacuum state ΩVS, the Bogoliubov implementer UV is formally defined on DF by

UVa(ϕ1)a(ϕn)Ω:=b(ϕ1)b(ϕn)ΩV, 79

with ϕDf and b(fj)=(a(ufj)+a(vfj¯)) for all basis vectors fj in f.

Lemma 5.3

(UV is well-defined) If ΩVSH^ (see (64)), then (79) defines an operator UV:DFS.

Proof

Both ufj and vfj¯ are proportional to the same basis vector ej (bosonic: gj, fermionic: ηj, see [45]). So the right-hand side of (79) is a finite linear combination of vectors a(ej1)a(ejn)ΩV. Now, ΩVS and by Lemma 4.9, each application of a(ej) leaves the vector in S.

Lemma 5.4

(Conditions for an implementer UV) Suppose that for a Bogoliubov transformation (i.e., V satisfying (24)) an ΩV satisfying b(ϕ)ΩV=0 for all ϕDf2 has been found, such that UV in (79) is well-defined on DF and has an inverse UV-1 defined on UV[DF]. Then, UV implements V in the sense of (77) on all ΨUV[DF].

Proof

We write Ψ=UVΦ with ΦDF. By linearity, it suffices to prove the statement for Φ=a(ϕ1)a(ϕn)Ω, which implies by (79) that Ψ=b(ϕ1)b(ϕn)ΩV. Checking the first statement of (77), i.e., UVa(ϕ)UV-1Ψ=b(ϕ)Ψ is straightforward. For the second statement, we make use of the CAR/CCR of a- and b-operators, using ε=(-1) for fermions and ε=1 for bosons. Here, the CAR/CCR are valid for a-operators by definition, and for b-operators, since by means of Lemma 4.1, the Bogoliubov relations survive the extension.

UVa(ϕ)UV-1Ψ==1nUVa(ϕ1)a(ϕ-1)ε+1ϕ,ϕa(ϕ+1)a(ϕn)Ω=b(ϕ)b(ϕ1)b(ϕn)ΩV-εn+1b(ϕ1)b(ϕn)b(ϕ)ΩV=(78)b(ϕ)b(ϕ1)b(ϕn)ΩV=b(ϕ)Ψ, 80

where we used the convention that the above sums are set to zero for N=0.

Bosonic Case

We now show that for a suitable choice of ΩV, the operator UV defined in (79) indeed implements the Bogoliubov transformation V.

Theorem 5.5

(Implementation via ITP works, bosonic) Consider a bosonic Bogoliubov transformation V=uvv¯u¯ with vv having countable spectrum. Let H^=kNHk be the ITP space (Definition 4.5) with respect to the basis (gk)kN2. Define the new vacuum vector

ΩV=kNΩk,V:=kN1-νk2μk21/4exp-νk2μk(a(gk))2Ωk, 81

where μk,νk are the singular values of uv as in Sect. 3.2. Then, V is implemented in the sense of (77) by UV:DFH^ (79).

Proof

By Lemma 5.4, we need to establish the following four points:

  1. The new vacuum ΩV is well-defined

  2. UV is well-defined on DF (Lemma 5.3 will be used, here)

  3. b(ϕ)ΩV=0

  4. UV-1 exists on UV[DF]

(1)Well-definednessofΩV_: Expression (81) is an ITP of one normalized factor per space Hk. Hence, it is a C-sequence, which can be identified with ΩVH^.

(2)Well-definednessofUV_: follows from Lemma 5.3, if we can establish ΩVS. By definition of S, we need to verify the rapid decay condition NknΩVck,nΩV. Since all Ωk,V are normalized, this boils down to proving NknΩk,Vk2ck,n2. We explicitly compute

NknΩk,Vk2=(1-4t2)1/2N=0t2N(2N)!(N!)2(2N)2n, 82

with t=νk2μk[0,1/2). Now, the function

Nt2N(2N)!(N!)2(2N)2n(2t)2N(2N)2n, 83

is positive, bounded and decays exponentially at N since 02t<1. So,

N=0t2N(2N)!(N!)2(2N)2ncons.+N=0(2t)2N(2N)2n=:ck,n2<, 84

which establishes ΩVS and hence the claim.

(3)b(ϕ)annihilatesΩV_: This is straightforward to check: Since ϕDf, the following sum over k is finite:

b(ϕ)ΩV=kϕk¯b(fk)ΩV. 85

As in the case, where the Shale condition holds, each b(fk) annihilates the corresponding vacuum vector Ωk,V, so the finite sum above is 0.

(4)Well-definednessofUV-1_: Consider the basis {a(fk1)a(fkN)ΩNN0,kN} of DF, where fk are chosen out of the basis (fj)jN (with j=k). If we can show that the set

{b(fk1)b(fkN)ΩVNN0,kN}H^, 86

with b(fk)=μka(gk)+νka(gk), is linearly independent, we are done, since then ker(UV)={0}, so UV is injective and hence invertible on its image.

Now, as applications of bk:=b(fk) and UV preserve the ITP structure, it suffices to show that on each mode k, the set

{(bk)NΩk,VNN0}Hk 87

is linearly independent. But (87) is just the image of the set

{(a(fk))NΩkNN0}F({fk}) 88

under a one-mode Bogoliubov transformation Uk,V:F({fk})Hk (defined as Uj,V in (33)). For a finite number m of modes, Bogoliubov transformations can always be implemented by unitary operators, as then the operator v:CmCm is always Hilbert–Schmidt. Now, (88) is an orthogonal set with no vector being 0, so its image (87) under Uk,V is also orthogonal with no vector being 0, and hence it is linearly independent.

Fermionic Case

Theorem 5.6

(Implementation via ITP works, fermionic) Consider a fermionic Bogoliubov transformation V=uvv¯u¯ with vv having countable spectrum. Let H^=kNHk be the ITP space (Definition 4.6). Define the new vacuum vector

ΩV=jJΩj,ViIΩ2i,2i-1,V:=jJ1a(ηj)ΩjjJ0ΩjiI(αi-βia(η2i)a(η2i-1))Ω2i,2i-1, 89

with αi,βi being the singular values of uv as in (40), and with Ω2i,2i-1,Ω2i,2i-1,VHk(i). Then, V is implemented in the sense of (77) by UV:DFH^ (79).

Proof

Again, by Lemma 5.4, it suffices to establish the four points in the proof of the bosonic case (Theorem 5.5). Points 1.) and 3.) are analogous, while 2.) follows from Lemma 5.3 as S=H^.

4.)Well-definednessofUV-1_: We proceed as in proof step 4.) in Theorem 5.5. So we are done if we can prove that the set

{b(fj1)b(fjN)ΩVNN0,jJ}H^ 90

is linearly independent. This again boils down to proving a linear independence statement on each Hk. The crucial difference now is, that each tensor product factor Hk may be a Fock space over either one or two modes. We abbreviate bj:=b(fj) and aj:=a(ηj). For two-mode factors indexed by iI, we need to prove linear independence of the set

{(b2i)N1(b2i-1)N2Ωk(i),VN1,N2{0,1}}Hk(i). 91

This follows from U2i,2i-1,V (see (42)) being unitary and mapping the set

{(a2i)N1(a2i-1)N2Ωk(i)N1,N2{0,1}}F({f2i})F({f2i-1}) 92

onto (91). For one-mode factors indexed by jJ we need linear independence of

{(bj)NΩk(j),VN{0,1}}Hk(j). 93

This follows again by unitarity of Uj,V, as well as orthogonality and zero-freeness of the set {(aj)NΩk(j)N{0,1}}F({fj}), which is mapped to (93). By linear independence of (91) and (93), we obtain linear independence of (90), which implies injectivity of UV and finishes the proof.

Remark 5.7

We may extend UV to a unitary operator on H^: Both in the bosonic and the fermionic case, we have H^=kNHk, compare (60) and (61), where a unitary Uk,V:HkHk is defined in (33) (bosonic) and (42) (fermionic). On tensor product states Ψ(m):=kΨk(m), we can thus immediately define UV as

UVΨ(m):=k(Uk,VΨk(m)),UVΨ(m)=Ψ(m). 94

By Lemma A.2, there exists an orthonormal set {Ψ(m)}mM such that any ΨH^ can be written as a convergent series Ψ=mMdmΨ(m), dmC. One easily checks that UV preserves orthonormality of Ψ(m). Thus, UV first extends to a bounded operator on all Ψ that are finite linear combinations of Ψ(m), and then to all ΨH^ by continuity. Unitarity of UV:H^H^ is then straightforward to check.

Remark 5.8

It is crucial that the fermionic ITP space has been chosen as H^=kHk, with two-mode spaces Hk=F({η2i})F({η2i-1}) for Cooper pairs iI. If we had just chosen a product of one-mode spaces jJF({ηj}), then invertibility of UV may fail.

As an example, consider a V with countably infinitely many Cooper pairs iI, such that αi=βi=12. Then, each Cooper pair is in the state

Ψi:=12(|0|0+|1|1)C4, 95

i.e., we have a “half particle–hole transformation”. When evaluating the formal ITP ΩV=iIΨi, we obtain a sum of C-sequences: For each pair i, one has to choose either |0|0 or |1|1 as a contribution to ΩV and sum over all choices. But now, there are uncountably many such choices, as each corresponds to a binary number of infinitely many digits. And each one gives a contribution of norm iI12=0. So ΩV=0, making UV non-invertible.

Diagonalization: Extended

As we now have conditions for V being implementable by UV in the extended sense, it would be interesting to diagonalize quadratic Hamiltonians H via UV. We now give precise definitions of “quadratic Hamiltonian” and “diagonalized” in the extended sense and provide diagonalizability criteria in Propositions 6.3 and 6.4.

Definition of Extended Diagonalization

Recall that the extended operator algebra A¯e, defined in (56), consists of all maps H that assign to each finite operator product a1am a complex coefficient Hj1,,jmC. Each map H can be interpreted as a (possibly infinite) sum

H=mj1,,jmHj1,,jma1am. 96

A formal quadratic Hamiltonian is an element HA¯e, where Hj1,,jm0 only appears for m=2 and H=H. We impose normal ordering on quadratic Hamiltonians (see Remark 6.2), so they read

H=12j,kN(2hjkajak±kjkajak+kjk¯ajak), 97

where ± means + in the bosonic and - in the fermionic case. The term “formal” stresses that H is not necessarily an operator on Fock space. To H we associate a block matrix

AH=h±kk¯±h¯, 98

with h=(hjk)j,kN, k=(kjk)j,kN being matrices of infinite size.

Consider a Bogoliubov transformation V=uvv¯u¯. Then, ajbj=k(ujkak+vjk¯ak), followed by a normal ordering, defines a corresponding algebraic Bogoliubov transformation VA¯:A¯edom(VA¯)A¯e, where dom(VA¯)A¯e is a suitable subspace that avoids diverging sums over k. The transformed operator and its associated block matrices are

H~=VA¯(H),AH~=VAHV. 99

Definition 6.1

A formal quadratic Hamiltonian HA¯e is called diagonalizable in the extended sense if there exists a Bogoliubov transformation V, such that

VAHV=E00±E, 100

with E0 being Hermitian, ± being + in the bosonic and - in the fermionic case, and where V is implementable in the extended sense (see Definition 5.1).

The Hamiltonian associated with AH~ is then H~=dΓ(E), where the matrix E provides a positive semidefinite quadratic form on De. So by Friedrichs’ theorem, it has a self-adjoint extension on dom(E). Following [48, Sect. VIII.10], dΓ(E) is then essentially self-adjoint on n=0dom(E)nF, so H~ defines quantum dynamics on F.

Remark 6.2

(Normal ordering constant) Our process of “diagonalizing” a Hamiltonian H actually consists of conjugating it with UV, so a is replaced by b, plus a subsequent normal ordering process. This process is equivalent to adding a constant to the Hamiltonian, namely

c=12tr(E)-tr(h)=12j(Ejj-hjj). 101

The sum might be divergent and hence not a complex number. Nevertheless, using ESS (see Appendix B), we may interpret it as an infinite renormalization constant cRen1(N), namely the one associated with the sequence cj=12(Ejj-hjj). If E now maps De into itself (so each column has finitely many non-zero entries), then

H~=UV-1(H+c)UV,

which is in accordance with (2). Otherwise, we may decompose H=nNH(n) and c=nNc(n), such that in VAH(n)V=E(n)00±E(n), each E(n) maps De into itself. So

H~=nNUV-1(H(n)+c(n))UV,

which is in accordance with (3).

Bosonic Case

Conditions for the existence of a V, such that VAHV is block-diagonal, can be found in [13, Thms. 1 and 4]. We can use them to readily derive conditions for when a formal quadratic Hamiltonian H is diagonalizable in the extended sense:

Proposition 6.3

(Extended diagonalizability, bosonic case) Let a formal quadratic bosonic Hamiltonian H (97) be given such that for the associated block matrix AH (98) we have h>0, and that G=h-1/2kh-1/2 is a bounded operator with G<1. Following [13, Theorem 1], there exists a bosonic Bogoliubov transformation V=uvv¯u¯ such that

VAHV=E00E. 102

Suppose further that vv has countable spectrum.

Then, H is diagonalizable in the extended sense.

Proof

Consider Definition 6.1 for diagonalizability. The existence of V as a block matrix associated with a bounded operator on 2 is a direct consequence of [13, Theorem 1].

If the spectrum of vv is countable, then implementability of V in the extended sense follows from Theorem 5.5.

Fermionic Case

Proposition 6.4

(Extended diagonalizability, fermionic case) Let a formal quadratic fermionic Hamiltonian H (97) be given such that for the associated block matrix AH (98), dimKer(AH) is even or . Following [13, Theorem 4], there exists some fermionic Bogoliubov transformation V=uvv¯u¯ such that

VAHV=E00-E. 103

Suppose further that vv has countable spectrum.

Then, H is diagonalizable in the extended sense on the ITP space H^.

Proof

Existence of a unitary V and of E0 follows from [13, Theorem 4]. By unitarity, VV=1=VV, so V is a fermionic Bogoliubov transformation. If σ(vv) is countable, then implementability follows from Theorem 5.6.

Applications

Quadratic Bosonic Interaction

Our first example for a quadratic Hamiltonian whose diagonalization requires Bogoliubov transformations “beyond the Shale/Shale–Stinespring condition” is inspired by [25]. We consider a free massive bosonic scalar field, which is interacting by a Wick square :ϕ(x)2:, with ϕ(x)=a(x)+a(x). We discretize the momentum by putting the system in a box x[-π,π]3 with periodic boundary conditions. Further, the Wick square is weighted by a real-valued external field κCc([-π,π]3),κ(x)C. The Hamiltonian then reads

H=dΓ(εp)+12κ(x):ϕ(x)2:dx=12p1,p2Z3(2εp1δ(p1-p2)a1ap2+2κ^(-p1+p2)a1ap2++κ^(-p1-p2)a1a2+κ^(p1+p2)ap1ap2), 104

with κ^(p)=κ^(-p)¯ denoting the Fourier transform of κ(x). For simplicity, we assume that κ(x)=const., so we can write κ^(p)=κδ(p), κR.

Proposition 7.1

For interactions κ>-m2 but κ0, the Hamiltonian H is diagonalizable in the extended sense on H^. However, the transformation V violates the Shale condition, so H is not diagonalizable on F.

Proof

We directly compute V and then apply Proposition 6.3. The matrix AH of A is

AH=pZ3AH,p,AH,p=hpkpkphp,hp=(εp+κ),kp=κ. 105

We diagonalize all AH,pC2×2 separately via

V=pZ3Vp,VpAH,pVp=Ep00Ep, 106

with Ep=h2-k2. Following [13, Sect. 1.3], this is done for |kp|<hp by

Vp=upvpvp¯up¯,up=cp,vp=cp-Gp1+1-Gp2, 107
Gp=kphp-1,cp=12+121-Gp2. 108

Now, |kp|<hp|κ|<|p|2+m2+κ, which is satisfied for all pZ3, if and only if κ>-m2. hp>0 also holds in that case and (106) defines a Bogoliubov transformation V diagonalizing AH. Regarding Proposition 6.3, h>0 and h-1/2kh-1/2<1 follow from hp and |kp|<hp after taking a direct sum. Since v=pZ3vp can be decomposed into modes, the same holds for vv, which therefore has countable spectrum. So Proposition 6.3 applies and H is diagonalizable in the extended sense on H^.

It remains to show that V violates the Shale condition, i.e., tr(vv)=pZ3|vp|2=. If |p| is large enough (say, |p|>pmax>0), we have κd|p|Gpκ|p| for any d<1, so

|vp|2=1+1-Gp221-Gp2Gp21+1-Gp22κ2d24|p|2. 109

We write p|vp|2 as an integral, using indicator functions χQ(p)(·) of half-open unit cubes Q(p) centered at p=(p1,p2,p3):

pZ3|vp|2=R3f(p)dp,f(p)=pZ3|vp|2χQ(p)(p). 110

Then, for |p|>pmax,

pZ3|vp|2|p|>pmaxf(p)dppmaxκ2d24(|p|+32)24π|p|2d|p|=, 111

where the integral is linearly divergent, which establishes the claim.

Remark 7.2

(Infinite volume case) Here, pR3 and we can construct an analogous V diagonalizing AH. However, the spectrum of vv is then no longer countable.

Remark 7.3

(Position-dependent κ(x)) In contrast to [25], we assume a constant interaction strength κ(x). Physically, it would be desirable to treat any κCc. Then, the decomposition AH=pZAH,p fails and it might occur that vv has uncountable spectrum.

Remark 7.4

(Wick square is not diagonalizable) It may be tempting to set εp=0 and to try a diagonalization of only the interaction Hamiltonian 12κ(x):ϕ(x):dx. However, bosonic Wick squares are not diagonalizable by a Bogoliubov transformation, as “the off-diagonal is too large”. For example, on one mode (hC), the matrix associated with a Wick square HW=2aa+aa+aa is

AHW=hkk¯h¯=1111, 112

so h-1/2kh-1/2=1 and Proposition 6.3 does not apply. However, bosonic Wick products have still been constructed as self-adjoint operators by a suitable GNS construction [49].

BCS Model

An example with non-implementable fermionic Bogoliubov transformations is the Bardeen–Cooper–Schrieffer (BCS) model for explaining superconductivity [23, 24]. For an overview on recent mathematical advances on BCS theory, we refer the reader to [50, 51] and the references therein. The “Hartree-like approximation” state [23, (2.16)] corresponds to a formal fermionic Bogoliubov vacuum state ΩV as in (89). A mathematical analysis by Haag [24] shows that in the infinite volume limit, the BCS Hamiltonian can indeed be diagonalized by a corresponding Bogoliubov transformation, which is not implementable on Fock space.

We consider a similar model of a fermionic gas inside a box with periodic boundary conditions x[-π,π]3. Hence, we have discretized momenta pZ3, as well as two spins s{,}, leading to a one-particle Hilbert space h=L2(Z3×{,}). The corresponding Fock space is F=F(Z3×{,}). We consider the following quadratic Hamiltonian (see [24]), which provides an approximate description for the fermionic gas that becomes exact in the infinite volume limit:

H=H0+HI=pZ3εpap,ap,+εpap,ap,-Δ~pap,ap,+Δ~p¯ap,ap,, 113

with kinetic energy εp=p22m-μR and interaction strength Δ~pC, of which we assume Δ~p0. As a basis (ej)jJ for identifying h with 2, we choose

(ep,s)pZ3s{,}L2(Z3×{,}),ep,s(p,s)=δppδss, 114

with δ being the Kronecker delta. The corresponding canonical basis of 2 is denoted (ep,s)pZ3,s{,}. In order to obtain momentum conservation, we have to interpret ap,,ap, as creating/annihilating a fermion of momentum -p. The mode index p is only used for an easier decomposition into modes.

Proposition 7.5

The Hamiltonian H (113) is diagonalizable in the extended sense on H^.

Proof

We compute V directly and apply Proposition 6.4. This is done in the block matrix representation AH=pAH,p and V=pVp with

AH,p=εp00-Δ~p0εpΔ~p00Δ~p¯-εp0-Δ~p¯00-εp,Vp=up00vp0up-vp00v¯pu¯p0-v¯p00u¯p, 115

where AH,p,Vp,sC4C4. The diagonalized matrix reads

VpAH,pVp=Ep0000Ep0000-Ep0000-Ep,Ep=εp2+|Δ~p|2, 116

and the diagonalization is established by

up=Δ~p(Ep-εp)2+|Δ~p|2,vp=Ep-εp(Ep-εp)2+|Δ~p|2. 117

From this form, it also follows that dimKer(AH) is either even or . So in order to apply Proposition 6.4, we only need to show that the spectrum of vv is countable. This is the case, since v=pZ3vp decays into modes, so also vv=pZ3(vv)p decays into modes, where each (vv)p is a finite-dimensional matrix with finite spectrum. As the sum over p is countable, also the spectrum of vv is countable, and by Proposition 6.4, H is diagonalizable on H^.

Remark 7.6

(Infinite volume case) The Hamiltonian H is an approximation to H=H0+HI, where HI is an attractive quartic interaction between fermion pairs. As mentioned above, this approximation is only exact in the infinite volume limit. In that case, pR3 and (115)–(117) still yield a Bogoliubov transformation V diagonalizing AH. However, vv has generally uncountable spectrum, so Theorem 5.6 does not apply.

Acknowledgements

I am grateful to Michał Wrochna, Roderich Tumulka, Andreas Deuchert, Jean-Bernard Bru and Niels Benedikter for helpful discussions. This work was financially supported by the DAAD (Deutscher Akademischer Austauschdienst) and also by the Basque Government through the BERC 2018-2021 program and by the Ministry of Science, Innovation and Universities: BCAM Severo Ochoa accreditation SEV-2017-0718, as well as by the European Union (ERC FermiMath, grant agreement nr. 101040991). Views and opinions expressed are however those of the author(s) only and do not necessarily reflect those of the European Union or the European Research Council Executive Agency. Neither the European Union nor the granting authority can be held responsible for them. Moreover, it was partially supported by Gruppo Nazionale per la Fisica Matematica in Italy.

Appendices

Results on the Infinite Tensor Product Space

Our first proposition characterizes when exactly two C-sequences correspond to the same functional ΨH^=kIHk. This allows us to define operators on H^ in Sect. 4.2. Recall that any C-sequence (Ψ)=(Ψk)kI gives rise to a unique functional ΨH^ by the embedding Ψ=ι((Ψ)).

Proposition A.1

Whenever (Ψ),(Ψ)Cseq represent the same functional Ψ=Ψ, then there exists a family of complex numbers (ck)kI such that

Ψk=ckΨkkI,andkIck=1, 118

using the notion of convergence for an infinite product from Sect. 2.2.

Conversely, if (Ψ),(Ψ)Cseq just differ by a family (ck)kI as in (118), then they represent the same functional Ψ=Ψ.

Proof

For the first statement, we must prove that Ψk and Ψk are parallel for any kI. So let us fix a k and decompose Ψk=Ψk+Ψk with ΨkΨk and ΨkΨk, and suppose that Ψk0. Now, choose some C-sequences (Φ),(Φ), that agree on all kk and with ΦkΨk and ΦkΨk, as well as Φkk=Φkk=1. Then,

Ψ,Φ=Ψk,ΦkkkkΨk,Φk=ΨkkkkΨk,Φkk<ΨkkkkΨk,Φkk=Ψk,ΦkkkkΨk,Φkk=Ψ,Φ. 119

By the same arguments, Ψ,Φ<Ψ,Φ. But since (Ψ),(Ψ) correspond to the same functional Ψ=Ψ, we can freely exchange both expressions within the scalar product:

Ψ,Φ=Ψ,Φ>Ψ,Φ=Ψ,Φ. 120

This contradicts (119) and thus establishes Ψk=ckΨk. Convergence of kIck can be seen as follows: We have

Ψ2=Ψ,Ψ=kIΨk,ckΨkk=kIckΨkk2. 121

So if kIck was not convergent, i.e., k|ck-1|=, then for the product on the right-hand side, we would have

kckΨkk2-1kΨkk2ck-1()-kΨkk2-1<. 122

Now, Ψkk2>1/2 for all but finitely many k, so () and thus the first expression in (122) diverges. This is a contradiction to (121) being convergent. So kIck indeed yields a complex number.

But since Ψ2=Ψ2=kIΨkk2, we immediately obtain kIck=1 from (121).

The converse statement can readily be seen by computing the action of the functionals Ψ,Ψ on some ΦCseq:

Ψ,Φ=kIΨk,Φkk=kIckΨk,Φkk=kIck¯kIΨk,Φkk=Ψ,Φ. 123

By [3, Lemma 4.1.1], all subspaces kICHk of H^=kIHk are mutually orthogonal. This allows for a particularly simple decomposition.

Lemma A.2

For any ΨH^, we can write

Ψ=mMdmΨ(m)=mMdmkIΨk(m), 124

with M being a subset of N, Ψ(m) defined by the mutually orthogonal C0-sequences (Ψ(m)) with Ψ(m)=1 and where m|dm|2< is a complex sequence.

Moreover, one can choose a fixed set Z={Ψ(a)}aA defined by mutually orthogonal, normalized C0-sequences (Ψ(a)), such that for all ΨH^, the form (124) can be achieved by taking only Ψ(m)Z. The decomposition (124) is then unique up to the choice of the Ψk(m) representing Ψ(m).

So Z is an orthonormal basis of H^ that might be uncountable, but the elements ΨH^ are all countable linear combinations with coefficient sequences in 2.

Proof

By definition, any ΨH^ can be approximated by a Cauchy sequence (Ψ(r))rN~kIHk,Ψ(r)Ψ. So each Ψ(r) can be written as a finite linear combination of C-sequences. All C-sequences, that are no C0-sequences, must have norm 0, so we drop them and simply write

Ψ(r)==1LrΨ(r), 125

with (Ψ(r)) being C0-sequences. Now, the C0-sequences decay into mutually orthogonal equivalence classes C, out of which countably many are occupied by any Ψ(r). So

Ψ(r)=C:(Ψ(r))CΨ(r)=:CΨC(r). 126

By orthogonality of the subspaces ΨkICHk, we have

Ψ(r)-Ψ(s)2=CΨC(r)-ΨC(s)2. 127

(Ψ(r))rN is a Cauchy sequence, so (ΨC(r))rN is also a Cauchy sequence for all C. That means, the limit ΨC=limrΨC(r) exists and by orthogonality of the ΨC(r) for each r,

limrCΨC(r)=ClimrΨC(r)Ψ=CΨC. 128

We may now write ΨC in coordinates:

ΨC=limrΨC(r)=limr:(Ψ(r))CΨ(r)=n(·)FaC(n(·))kIek,n(k), 129

so also Ψ can be written as a countable sum over mutually orthogonal, normalized C0-sequences with coordinates aC(n(·)). We index the sequences and coordinates by (Ψ(m)) and dm, which yields the desired form (124).

Square summability of the dm can be seen by

m|dm|2=Cn(·)F|aC(n(·))|2. 130

Now, the set Z={Ψ(a)}aA is exactly the union of all vectors kIek,n(k) over all classes C, which is indeed a set of mutually orthogonal, normalized vectors.

Uniqueness of the decomposition follows by orthogonality of the Ψ(r).

Further, by [3, Definition 6.1.1], two C0-sequences (Ψ),(Φ) are weakly equivalent, if and only if there exists a family (zk)kIC with (zkΨk)kI being (strongly) equivalent to (Φk)kI. From that, we conclude:

Lemma A.3

Let Cw be the weak equivalence class of a C0-sequence (Φ)=(Φk)kI, choose an orthonormal basis (ek,n)nN0 for each Hk such that Φk=ck,0 and define for any C0-sequence (Ψ)=(Ψk)kI the coordinates ck,n:=ek,n,Ψkk.

Then, kICwHk is exactly the closure of the span of all normalized C0-sequences, where |ck,0|=1 for all but finitely many kI.

Note that the last statement means ck,n=0 for n1 and for those k. In simple words, Lemma A.3 asserts that replacing C by Cw in the equivalence class is done via replacing ck,0=1 by |ck,0|=1.

Proof

First, we prove that any normalized C0-sequence (Ψ) with |ck,0|=1 almost everywhere is weakly equivalent to (Φ): We can define a family of phase rotations |zk|=1, such that zkck,0=1 for all k with |ck,0|=1. So (zkΨk)kI has zkck,0=1 almost everywhere and is hence a C0-sequence strongly equivalent to (Φ). Therefore, (Ψ) is weakly equivalent to (Φ). So ΨkICwHk and the same holds for the span of these C0-sequences and their closure with respect to the Hilbert space topology on H^.

Conversely, any ΨkICwHk is within the closure of the span of normalized C0-sequences with |ck,0|=1 almost everywhere: By Lemma A.2, we may write

Ψ=mMdmkIΨk(m),

where m|dm|2< and the (Ψ(m))=(Ψk(m))kI with Ψ(m)=1 are orthogonal. Further, we may choose (Ψ(m))w(Φ), since all (Ψ(m)) were constructed to come from a (strong) equivalence class C contained within Cw. So there exist families (zk(m))kI,|zk(m)|=1, such that (zk(m)Ψk(m))kI(Φ) for all mM. By strong equivalence, we may approximate each (zk(m)Ψk(m)) up to arbitrary precision ε>0 by a linear combination of (normalized) families Ψ(m,ε), such that, when writing these families in coordinates, we have ck,0(m,ε)=1 almost everywhere in kI. Hence, the families ((zk(m))-1Ψk(m,ε))kI approximate Ψ(m) up to precision ε. They satisfy |(zk(m))-1ck,0(m,ε)|=1 almost everywhere, so Ψ(m) can be approximated up to arbitrary precision by a linear combination of C0-sequences with the above-mentioned property. And by (124) and convergence of |dm|2, also an arbitrary approximation of Ψ is possible by linear combinations of normalized C0-sequences with |ck,0|=1 almost everywhere.

Extended State Space

As mentioned in the introduction, it is also possible to achieve results similar to Theorems 5.5 and 5.6 in the “Extended State Space” setting from [26, 27]. In this appendix, we define extensions F¯,F¯ex adapted to vv having discrete spectrum, so the definitions are simpler than in [26, 27].

ESS Construction

We start by defining

  • The space of generalized one-particle wave functions
    E=E(N)=CN:={ϕ:NC}, 131
    that is, the space of complex sequences ϕ=(ϕj)jN. It extends Dg (bosonic) or Dη (fermionic) and replaces the smooth function space S˙1 from [26].
  • The space of generalized N-particle wave functions
    E(N)(N):={Ψ:NNC}. 132
  • The space of generalized Fock space functions (which replaces S˙F from [26]):
    EF(N):=N=0E(N)(N)={Ψ:Q(N)C}. 133

EF is not yet the final extended state space. Formal state vectors occurring in QFT include products of functions ΨmEF and exponentials of divergent sums er with “r=±”, see [26]. Hence, we need to construct structures accommodating such infinite quantities.

First, we introduce a space of renormalization factors Ren1(N), whose elements are sequences that represent formal (and possibly divergent) series, thus replacing the formal (and possibly divergent) integrals from Ren1 in [26].

  • Ren1(N):=E/Ren1. Here, for r1,r2E(N), we define r1Ren1r2 if and only if (r1-r2)1=L1(N) and jN(r1,j-r2,j)=0.

We denote elements of Ren1 by r=[r] and identify r=jNrjC, if r1. All Ren1-elements not identifiable with a C-number can be thought of as “controlled infinitely large numbers”.

Multiplication of Ren1-elements is allowed by introducing the free vector spaces PolP(N), PN, called space of renormalization polynomials of degree P, and defined as being spanned by all commutative products r1··rp,pP,rjRen1(N). Again, we identify equivalent terms by modding out an equivalence relation:

  • RenP(N):=PolP(N)/RenP with RenP generated by both r1r2rpRenPc1r2rp for r1=jNr1,j=c1C and (c1c2)r1rpRenPc1(c2r1)rp.

The space of renormalization polynomials is then given by

Ren(N):=PNRenP(N). 134

Wave function renormalization factors usually take the form er, where rRen1 may or may not correspond to a finite number, as well as linear combinations of such terms er. We will interpret them as elements of a suitably defined field eRen. For this, consider the group algebra C[Ren1], where the group is (Ren1,+) with addition as described above. We identify elements rRen1 with the symbolic expression er, both to distinguish the two additions in (Ren1,+) and C[Ren1], as well as to be congruent with the intuition that er is an exponential of a finite or infinite number. Elements in C[Ren1] are then denoted as c1er1++cMerM,cjC,rjRen1. Some elements of C[Ren1] are intuitively equal to zero. We set them equivalent to zero by modding out an ideal IC[Ren1] generated by all elements ecer-ec+r with cC,rRen1.

As in [26, Proposition 3.2] one proves that the quotient ring C[Ren1]/I has no proper zero divisors. So the following quotient field exists, which is a field extension of C:

  • eRen(N):={c=a1/a2a1,a2C[Ren1(N)]/I} is the field of wave function renormalizations.

The formal state vectors appearing in QFT are of the form Ψ=mcmΨm with cmeRen and ΨmEF. Those can be described as elements of the following space:

  • F¯(N):=F¯0(N)/F is the first extended state space, where F¯0(N) is the free eRen-vector space over EF (all finite eRen-linear combinations) and F is generated by (cc)ΨmFc(cΨm) for cC.

For intermediate calculations, we will need an even larger vector space F¯ex, which allows for multiplication of Ψ by elements of Ren. We define RenQ(N) to consist of all functions Q(N)Ren, which replaces the similarly-defined RenQ˙ from [26].

  • F¯ex(N):=F¯ex,0(N)/Fex is the second extended state space, where F¯ex,0(N) is the set of all countable eRen-linear combinations Ψ=mNcmΨm with cmeRen(N),ΨmRenQ(N) and where Fex is generated by (cc)ΨmFexc(cΨm) for cC.

We may embed the complex numbers into Ren by identifying cC with ce0eRen. Hence, the space F¯ can be embedded into F¯ex.

Remark B.1

(Comparison with non-standard analysis) The construction of Ren1 might remind about that of the extended real line * R in non-standard analysis [52, Sect. 1.1], since both extend R by modding out an equivalence relation on sequences. One may therefore ask whether Ren1 can be naturally embedded into * R or vice versa. This is not the case. First, note that Ren1 extends C and not just R, while there are no elements corresponding to complex numbers in * R. Even when restricting to real sequences, an embedding would require to identify a sequence r=(rj)jN, [r]Ren1, with its series s(r):=(s(r)j)jN, s(r)j:==1jr with respect to * R. This is needed to be consistent with the identification r=cR in Ren1 whenever jrj=c<, as well as the identification s(r)=cR in * R, whenever s(r)j=c for all j>J, JN.

Now, there always exists an infinite set AN, such that s(r1) and s(r2) equivalent with respect to * R whenever they agree on N\A. However, the difference between r1 and r2 on A may not be 1-summable, rendering the two sequences inequivalent with respect to Ren1. So Ren1 cannot be naturally embedded into * R.

Conversely, * R contains many inequivalent s(r) converging to the same cR (corresponding to infinitesimals), even with 1-summable differences, while the corresponding r are all equivalent with respect to Ren1.

Operator Lift and Implementation

Creation and annihilation operators a(ϕ),a(ϕ) are defined on Ψm:Q(N)C in similarity to (10). Formally,

(a±(ϕ)Ψm)(q)=k=1N(±1)kNϕjkΨm(q\jk),(a±(ϕ)Ψm)(q)=N+1jϕj¯Ψm(q,j). 135

Using that there is a natural distribution pairing between D and E, one easily verifies the following well-definedness statement.

Lemma B.2

(Products of a,a are well-defined on the ESS) Consider the ESS F¯ built over E. Then, (135) uniquely defines operators a(ϕ),a(ϕ) as follows:

a(ϕ):F¯F¯,a(ϕ):F¯F¯ϕD,a(ϕ):F¯F¯,a(ϕ):F¯F¯exϕE. 136

Note that the CAR/CCR are a direct consequence of (135) and hence still valid for the operator extensions.

The condition for UV:DFF¯ being an “extended implementer” is identical to the ITP case (Definition 5.1). Also Definition 5.2 for UV, given a Bogoliubov vacuum ΩVF¯, carries over from ITP to ESS. Then, Lemma 5.3 (well-definedness of UV) and Lemma 5.4 are established analogously. By checking the four conditions as in the proofs of Theorems 5.5 and 5.6, it is then straightforward to verify the following two results.

Proposition B.3

(Implementation via ESS works, bosonic) Consider a bosonic Bogoliubov transformation V=uvv¯u¯ with vv having countable spectrum. Let F¯ be the ESS over E with respect to the basis (gk)kN2. Define the new vacuum vector

ΩV=exp14klog1-νk2μk2=:erexp-kνk2μk(a(gk))2=:ΨVΩ=erΨV, 137

where μk,νk are the singular values of uv as in Sect. 3.2. Then, V is implemented in the sense of (77) by UV:DFF¯ (79).

Proposition B.4

(Implementation via ESS works, fermionic) Consider a fermionic Bogoliubov transformation V=uvv¯u¯ with vv having countable spectrum. Let F¯ be the ESS over E with respect to the basis (ηj)jJ, and let |J1|<, so the number of modes with a full particle–hole transformation is finite. Define the new vacuum vector

ΩV=expiIlogαi=:erjJ1a(ηj)iI1-βiαia(η2i)a(η2i-1)=:ΨVΩ=erΨV, 138

with αi,βi being the singular values of uv as in (40). Then, V is implemented in the sense of (77) by UV:DFF¯ (79).

In contrast to Theorem 5.6, the additional condition of having finitely many modes with a particle–hole transformation comes from the requirement that ΩV be in F¯.

Diagonalizability of Hamiltonians in an extended sense can then be defined for F¯ exactly as for H^ (Definition 6.1). Finally, also Propositions 6.3 and 6.4 for diagonalizability carry over from H^ to F¯, with the restriction to finitely many particle–hole transformed modes in the fermionic case.

Funding

Open access funding provided by Università degli Studi di Milano within the CRUI-CARE Agreement.

Data Availability

No datasets were generated during this work.

Declarations

Conflict of interest

The author has no conflicts to disclose.

Footnotes

1

In relativistic QFT, Haag’s theorem forbids a common representation for free and interacting dynamics [43, Sect. II.1] But such a common representation may also fail to exist in more general situations, like in polaron models [33] or when taking the thermodynamic limit of many-body systems [10, Chapter 17].

2

Here, for A-, the term “generated” comprises taking the closure.

3

See also the formulation [47, Thm. 10.10] of the spectral theorem.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  • 1.Dedushenko, M.: Snowmass white paper: the quest to define QFT int. J. Mod. Phys. A 38(4), 2330002 (2023) [Google Scholar]
  • 2.Summers, S.J.: A perspective on constructive quantum field theory. arXiv:1203.3991v2
  • 3.Neumann, Jv.: On infinite direct products. Comput. Math. 6, 1–77 (1939) [Google Scholar]
  • 4.Chung, V.: Infrared divergence in quantum electrodynamics. Phys. Rev. 140, B1110 (1965) [Google Scholar]
  • 5.Kibble, T.W.B.: Coherent soft-photon states and infrared divergences. I. Classical currents. J. Math. Phys. 9, 315 (1968) [Google Scholar]
  • 6.Kulish, P.P., Faddeev, L.D.: Asymptotic conditions and infrared divergences in quantum electrodynamics. Theor. Math. Phys. 4, 745–757 (1970) [Google Scholar]
  • 7.Blanchard, P.: Discussion mathématique du modè1e de Pauli et Fierz relatif à la catastrophe infrarouge. Commun. Math. Phys. 15, 156–172 (1969) [Google Scholar]
  • 8.Fröhlich, J.: On the infrared problem in a model of scalar electrons and massless, scalar bosons. Ann. Henri Poincaré 19, 1–103 (1973) [Google Scholar]
  • 9.Könenberg, M., Matte, O.: The mass shell in the semi-relativistic Pauli–Fierz model. Ann. Henri Poincaré 15, 863–915 (2014) [Google Scholar]
  • 10.Dereziński, J., Gérard, C.: Mathematics of Quantization and Quantum Fields. Cambridge University Press, Cambridge (2013) [Google Scholar]
  • 11.Araki, H.: On the diagonalization of a bilinear Hamiltonian by a Bogoliubov transformation. Publ. Res. Inst. Math. Sci. 4(2), 387–412 (1968) [Google Scholar]
  • 12.Bach, V., Bru, J.B.: Diagonalizing quadratic bosonic operators by non-autonomous flow equation. Mem. Am. Math. Soc. 240, 1138 (2016) [Google Scholar]
  • 13.Nam, P.T., Napiórkowsi, M., Solovej, J.P.: Diagonalization of bosonic quadratic Hamiltonians by Bogoliubov transformations. J. Funct. Anal. 270(11), 4340–4368 (2016) [Google Scholar]
  • 14.Matsuzawa, Y., Sasaki, I., Usami, K.: Explicit diagonalization of pair interaction models. Anal. Math. Phys. 11, 48 (2021) [Google Scholar]
  • 15.Schlein, B.: Bogoliubov excitation spectrum for Bose-Einstein condensates. In: Proceedings of the ICM 2018, pp. 2669–2686. (2019) arXiv:1802.06662v1
  • 16.Giacomelli, E.L.: Bogoliubov theory for the dilute Fermi gas in three dimensions. In: Proceedings of the Intensive Period “INdAM Quantum Meetings (IQM22)” at Politecnico di Milano (2023). arXiv:2207.13618
  • 17.Giacomelli, E.L.: An optimal upper bound for the dilute Fermi gas in three dimensions. J. Funct. Anal. 285(8), 110073 (2023) [Google Scholar]
  • 18.Falconi, M., Giacomelli, E.L., Hainzl, C., Porta, M.: The dilute Fermi gas via Bogoliubov theory. Ann. Henri Poincaré 22, 2283–2353 (2021) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Shale, D.: Linear symmetries of free Boson fields. Trans. Am. Math. Soc. 103, 149–167 (1962) [Google Scholar]
  • 20.Ruijsenaars, S.N.M.: On Bogoliubov transformations for systems of relativistic charged particles. J. Math. Phys. 18, 517 (1977) [Google Scholar]
  • 21.Torre, C.G., Varadarajan, M.: Functional evolution of free quantum fields. Class. Quantum Gravity 16, 2651–2668 (1999) [Google Scholar]
  • 22.Marecki, P.: Quantum electrodynamics on background external fields. PhD Thesis, University of Hamburg (2003). https://arxiv.org/pdf/hep-th/0312304.pdf
  • 23.Bardeen, J., Cooper, L.N., Schrieffer, J.R.: Theory of superconductivity. Phys. Rev. 108, 1175 (1957) [Google Scholar]
  • 24.Haag, R.: The mathematical structure of the Bardeen–Cooper–Schrieffer model. Nuovo Cimento 25(2), 287–299 (1962) [Google Scholar]
  • 25.Derezińki, J.: Bosonic quadratic Hamiltonians. J. Math. Phys. 58, 121101 (2017) [Google Scholar]
  • 26.Lill, S.: Extended state space for describing renormalized Fock spaces in QFT. Rev. Math. Phys. 36(8), 2450029 (2024) [Google Scholar]
  • 27.Lill, S.: Bogoliubov transformations beyond Shale–Stinespring: generic Inline graphic for Bosons, In: Proceedings of the Intensive Period “INdAM Quantum Meetings (IQM22)” (2023). arXiv:2208.03487
  • 28.Nelson, E.: Interaction of nonrelativistic particles with a quantized scalar field. J. Math. Phys. 5, 1190 (1964) [Google Scholar]
  • 29.Glimm, J.: Boson fields with the: Inline graphic: interaction in three dimensions. Commun. Math. Phys. 10, 1–47 (1968) [Google Scholar]
  • 30.Dereziński, J.: Van Hove Hamiltonians—exactly solvable models of the infrared and ultraviolet problem. Ann. Henri Poincaré 4, 713–738 (2003) [Google Scholar]
  • 31.Glimm, J., Jaffe, A.: The Inline graphic quantum field theory without cutoffs. I. Phys. Rev. 176, 1945 (1968) [Google Scholar]
  • 32.Glimm, J., Jaffe, A.: Self-adjointness of the Inline graphic, Hamiltonian. Ann. Phys. 60, 321–383 (1970) [Google Scholar]
  • 33.Gross, L.: The relativistic polaron without cutoffs. Commun. Math. Phys. 31, 25–73 (1973) [Google Scholar]
  • 34.Teufel, S., Tumulka, R.: Avoiding ultraviolet divergence by means of interior-boundary conditions. In: Finster, F., et al. (eds.) Quantum Mathematical Physics, pp. 293–311. Springer, Cham (2016) [Google Scholar]
  • 35.Lampart, J., Schmidt, J., Teufel, S., Tumulka, R.: Particle creation at a point source by means of interior-boundary conditions. Math. Phys. Anal. Geom. 21, 12 (2018) [Google Scholar]
  • 36.Lampart, J., Schmidt, J.: On Nelson-type Hamiltonians and abstract boundary conditions. Commun. Math. Phys. 367, 629–663 (2019) [Google Scholar]
  • 37.Lampart, J.: A nonrelativistic quantum field theory with point interactions in three dimensions. Ann. Henri Poincaré 20, 3509–3541 (2019) [Google Scholar]
  • 38.Schmidt, J.: On a direct description of pseudorelativistic Nelson Hamiltonians. J. Math. Phys. 60, 102303 (2019) [Google Scholar]
  • 39.Lienert, M., Nickel, L.: Multi-time formulation of particle creation and annihilation via interior-boundary conditions. Rev. Math. Phys. 32(2), 2050004 (2020) [Google Scholar]
  • 40.Henheik, J., Tumulka, R.: Interior-boundary conditions for the dirac equation at point sources in 3 dimensions. J. Math. Phys. 63, 122302 (2020) [Google Scholar]
  • 41.Lampart, J.: The renormalised Bogoliubov–Fröhlich Hamiltonian. J. Math. Phys. 61, 101902 (2020) [Google Scholar]
  • 42.Schmidt, J.: The massless Nelson Hamiltonian and its domain. In: Michelangeli, A. (ed.) Mathematical Challenges of Zero-Range Physics, pp. 57–80. Springer, Cham (2021) [Google Scholar]
  • 43.Haag, R.: Local Quantum Physics. Springer, Cham (1996) [Google Scholar]
  • 44.Berezansky, Yu. M., Kondratiev, Y.G.: Spectral Methods in Infinite-Dimensional Analysis. Springer, Cham (1995) [Google Scholar]
  • 45.Solovej, J.P.: Many body quantum mechanics lecture notes. http://web.math.ku.dk/~solovej/MANYBODY/mbnotes-ptn-5-3-14.pdf (2014) Accessed 1 July 2021
  • 46.Bach, V., Lieb, E.H., Solovej, J.P.: Generalized Hartree–Fock theory and the Hubbard model. J. Stat. Phys. 76, 3–89 (1994) [Google Scholar]
  • 47.Hall, B.C.: Quantum Theory for Mathematicians. Springer, Cham (2013) [Google Scholar]
  • 48.Reed, M., Simon, B.: Methods of Modern Mathematical Physics I: Functional Analysis. Academic Press, San Diego (1980) [Google Scholar]
  • 49.Sanders, K.: Essential self-adjointness of Wick squares in quasi-free Hadamard representations on curved spacetimes. J. Math. Phys. 53, 042502 (2012) [Google Scholar]
  • 50.Hainzl, C., Seiringer, R.: The Bardeen–Cooper–Schrieffer functional of superconductivity and its mathematical properties. J. Math. Phys. 57, 021101 (2016) [Google Scholar]
  • 51.Henheik, J., Lauritsen, A.B., Roos, B.: Universality in low-dimensional BCS theory. Rev. Math. Phys. 36(9), 2360005 (2024) [Google Scholar]
  • 52.Albeverio, S., Fenstad, J.E., Høegh-Krohn, R., Lindstrøm, T.: Nonstandard Methods in Stochastic Analysis and Mathematical Physics. Academic Press, San Diego (1986) [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

No datasets were generated during this work.


Articles from Journal of Statistical Physics are provided here courtesy of Springer

RESOURCES