Bogoliubov Transformations Beyond Shale–Stinespring: Generic for bosons
Abstract
We construct an extension of Fock space and prove that it allows for implementing bosonic Bogoliubov transformations in a certain extended sense. While an implementation in the regular sense on Fock space is only possible if a certain operator is trace class (this is the well–known Shale–Stinespring condition), the extended implementation works without any restrictions on this operator. This generalizes a recent result of extended implementability, which required to have discrete spectrum.
1 Introduction
Bogoliubov transformations are a versatile tool for elucidating the physical properties of many–body and quantum field theory (QFT) models in mathematical physics. Applications include the simplification of Hamiltonians in interacting bosonic or fermionic –body systems [1, 2, 3, 4, 5], QFT toy models [6, 7, 8] and relativistic quantum dynamics [9, 10].
In simple words, a Bogoliubov transformation is a linear replacement of creation– and annihilation operators such that satisfy the same commutation relations as . There are cases where allows for finding a unitary operator on Fock space such that and . In this case, is called an implementer of and is called implementable. The case of implementability is interesting for the following reason: Suppose, we are given a Hamiltonian , which is an inconvenient sum of products of , but there exists a Bogoliubov transformation and a constant , such that takes a much more convenient form in the –operators. This situation arises frequently for quadratic Hamiltonians, see Remark 5. Then, , which is implemented by , replaces by . So, under the replacement, becomes
(1) |
which has a convenient form in the –operators. This convenient form allows for a simple description of the dynamics generated by , which is by (1) unitarily equivalent to the dynamics generated by .
The important question when is implementable has long been settled by Shale and Stinespring [11, 12]: This is the case if and only if for some operator (16), characterizing , we have
This so–called Shale–Stinespring condition is rather restrictive. Therefore, the author recently [13] proposed the construction of an implementer in an extended sense, , which also achieves the replacement induced by , but can be constructed for a much greater class of Bogoliubov transformations. In [13], two Fock space extensions were constructed, each together with an operator or , with being a dense subspace, such that (1) holds as a strong operator identity on . Here, is an infinite tensor product space of the form introduced by von Neumann [14] and is an “extended state space” as recently introduced in [15]. The main result in [13] was that can be implemented in the extended sense, whenever
In this general case, becomes an “infinite renormalization constant”, which was rigorously interpreted in [13] as an element of some vector space .
The purpose of this article is to propose a new Fock space extension , related to , which allows for an extended implementer
as long as is defined on a suitable domain. We establish the notation in Section 2, construct in Section 3, state our main result, Theorem 4.1, in Section 4 and prove it in Section 5.
An analogous result can be expected to hold for fermions, as long as 1 is not an eigenvalue of infinite multiplicity of , i.e., only performs a particle–hole transformation on finitely many modes. The main technical complication in the fermionic case comes from the fact that the operator (47) is unbounded, whereas it is bounded in the bosonic case.
An extended implementer is particularly useful in cases where is an algebraic expression that does not define an operator on a dense subspace of , while maps into itself and allows for a self–adjoint extension. In that case, we can mathematically make sense of as an operator mapping into itself (see Figure 1) and use the dynamics generated by on for a physical interpretation.
The construction of is significantly shorter than that of in [13]. We intend a future use of as a bookkeeping tool for more general operator transformations that go beyond Bogoliubov transformations.
Right: Sketch of an extended implementation using the Fock space extension .
2 Notation
The notation is adapted from [13] in great parts. We consider a system with an indeterminate number of particles , whose one–particle sector is given by the sequence space .333Note that any separable Hilbert space can be identified with by fixing an orthonormal basis, so the description with is quite general. The configuration of the system is given by a vector of mode numbers , which is an element of configuration space444The symbol denotes the disjoint union of two sets. This notation is chosen to emphasize that, e.g., is not just .
(2) |
Here, is also called the –particle sector of configuration space. The state of the system is described by a normed vector in the (mode–) Fock space
(3) |
We impose bosonic symmetry using the symmetrization operator with
(4) |
where is the permutation group. The bosonic Fock space is then given by
(5) |
Equivalently, using the symmetric tensor product
(6) |
we can write
(7) |
Denote by the canonical basis of and by the creation and annihilation operators corresponding to , which are a priori just symbolic expressions. Products of these expressions or countable sums thereof are not necessarily defined as Fock space operators. Instead, we interpret them as elements of the ∗–algebra , defined as follows: Denote by
(8) |
the set of all finite operator products with respect to the basis . Then, is defined as the set of all countable sums
(9) |
with depending on . Let us denote the space of all complex–valued sequences by
(10) |
For , we then introduce the creation/annihilation operators
(11) |
where the overline denotes complex conjugation. Alternatively, complex conjugation is described by the complex conjugation operator
(12) |
with and whose adjoint satisfies . We will interchangeably use the notations and for elements of a series . Further, for we impose the canonical commutation relations (CCR) on
(13) |
In case , we may also densely define as Fock space operators
(14) | ||||
which preserve symmetry and satisfy (13) as a strong operator identity on a dense domain in Fock space.
Creation and annihilation operators are particularly easy to handle, if the sum over is finite, or equivalently, if the form factor is an element of the following space:
(15) |
Note that and , i.e., is the dual space of with respect to a suitable seminorm–induced topology, see [16, Chap. 1, Example 1.6].
By an algebraic Bogoliubov transformation, we mean a map , that sends
(16) | ||||
for all , where satisfy the bosonic Bogoliubov relations as a weak operator identity:
(17) | |||||
Here, is the adjoint of and the conjugate and transposed operators are given by and (and the same for ). Writing as matrices of infinite size, , we equivalently have , and . Further, we may keep track of by a block matrix . In [13, Lemma 4.1] it was proven that (17) holds, whenever and both preserve the CCR, that is, (13) also holds for replaced by .
3 Construction of the Fock space extension
We will establish implementability on a Fock space extension here, which is related, but not equal to the extensions in [13] and [15]. Just as in [13], we define the generalized –particle space and the generalized Fock space as
(18) |
The definition of is motivated by formal and possibly divergent sums that appear when formally applying one or several operators as in (14) to functions in . In its most general form, such a formal sum on the –sector reads
(19) |
where . There are two ways to read (19): First, can be viewed as a generalized function value, which is given by a formal sum at fixed . Second, (19) can be seen as a definition for a generalized function on that is obtained by taking the “raw functions” and formally “integrating out” the last variables.
Mathematically, the sector is specified by a pair . In order to allow for countable linear combinations and products of such expressions, we introduce the algebra of functions
(20) |
Here, is to be understood as a –vector space with linear combinations defined argument–wise, i.e.,
(21) |
The algebra multiplication is given by
(22) |
with denoting the topological tensor product in . We now introduce the sum notation, where we write each function in as a formal countable sum
(23) |
This notation is consistent with the notation in (19), and (22) coincides with the formal multiplication of two sums of the kind (23). However, permuting sum indices or executing convergent sums in (23) intuitively leaves the sum invariant, but results in a different associated element in . Therefore, we mod out a (two–sided) ideal555By an ideal of an algebra, we mean that shall be closed (in an algebraic sense) under the (commutative) algebra multiplication and linearity. So for and , we have and . Intuitively, the elements in are the ones equivalent to 0. , which is generated by requiring that:
-
(A)
, whenever each can be obtained by permuting the last indices of . So we may swap indices that are integrated out.
-
(B)
, whenever we may choose for each pair some integer , such that
(24) is an absolutely convergent sum for all and all . In other words, we may execute absolutely convergent sums, where is the number of sums executed in .
The extended state space is then defined as the quotient algebra
(25) |
Its elements (which are cosets of ) can thus be treated as if they were formal sums (23) and we adopt the sum notation also for elements of .
Remarks.
-
1.
extends : In sum notation (23), each corresponds to exactly one element of , viz. the one sending and for .
Further, if with , then the associated element in is not in . This is because (23) renders a complex number at each fixed , which is nonzero for at least one and cannot be changed by executing/un–executing convergent sums or permuting sum indices. So each can be identified with a distinct coset of , i.e., with an element of . So we may embed into and thus, is a true Fock space extension. -
2.
Tensor products on : The –product on (22) induces an –product on the quotient algebra . With the tensor product , one can also view as a “double–graded algebra” with degrees , just as together with the tensor product becomes a graded algebra (with degree ). Also, is a graded algebra with degree , but not “double–graded”, since is not unique.
-
3.
Comparison with [13]: The construction of deviates from the construction of the Fock space extensions in [13]. In particular, the latter construction renders some additional vector spaces as a byproduct. Let us quickly comment on how to find certain elements of these spaces in :
Elements are a rigorous implementation of formal sums
(26) for some fixed . Mathematically, is an equivalence class with
(27) where the sum is required to be absolutely convergent. It is easy to see that (26) is a sum notation of the coset666Here, denotes the coset of in , which contains , and is the Kronecker delta. , so the pair is mapped to and all other are mapped to 0. Further, for , we have if and only if (27) holds, since
(28) by definition of addition in .
Elements are now linear combinations of products of the form . These products rigorously implement formal sums
(29) with being a representative function of . It is easy to see that in the sum notation.
Finally, elements can be seen as a rigorous implementation of
(30) Obviously, . However, generic elements of are quotients of the form with being a linear combination of terms with . Such a quotient can generally not be written as a sum of the form (19). So we cannot readily find in and thus also not the –vector spaces and .
4 Main Result
Consider any Bogoliubov transformation with (recall: in (15) is the set of all sequences with finite support), as well as the dense domain
(31) |
where , defined by and for , is the vacuum vector. In analogy to [13, Def. 5.1], we define extended implementability as follows.
Definition 4.1.
A linear injective operator implements a Bogoliubov transformation in the extended sense if and only if
(32) |
In that case, is called an extended implementer of and is called implementable in the extended sense.
Our main result is the following.
Theorem 4.1 (Bosonic Extended Implementation Works).
Any bosonic Bogoliubov transformation with is implementable in the extended sense.
Remarks.
-
4.
Regular implementability: Definition 32 generalizes the (regular) notion of implementability of on Fock space. In the regular sense, is implementable if and only if there exists some unitary satisfying (32). By taking limits in , the case can then be generalized to .
While regular implementability only holds for [11, 12], extended implementability can be achieved for any , as long as is defined on . -
5.
Diagonalization: Extended implementers can be used to diagonalize quadratic Hamiltonians in some extended sense, as described in [13, Sect. 6]: Consider a diagonalizable quadratic Hamiltonian , that is,
(33) In certain cases [17, 18], there is an algebraic Bogoliubov transformation and a normal ordering constant , such that becomes diagonal in the –operators, so is block–diagonal in the –operators:
(34) and such that is self–adjoint on some domain . In that case, is also self–adjoint. Further, if maps , then
(35) holds as a strong operator identity . So is diagonalized in the extended sense by . Here is a possibly divergent sum of the form (26), so we can interpret , and a multiplication by maps .
5 Proof of the Main Result
5.1 General Construction of the Extended Implementer
We now construct the extended implementer for a given Bogoliubov transformation . The construction employs creation and annihilation operators , which we first need to extend to in such a way that they satisfy the CCR. Consider an element characterized by a coset element (“representative”) . In analogy to (14), we then set
(36) | ||||
with and .
Lemma 5.1 (Operator Extensions).
For , the representative–wise definition (36) renders well–defined linear operators
(37) |
Proof.
Consider the two expressions in (36). It is clear that the functions are elements of . So it remains to show that for , we also have .
First, suppose that where is obtained from by a permutation of the last indices (i.e., those in ). This situation corresponds to case (A) above (25). Then the same index permutation in transforms into . Thus, . A similar argument shows that when is obtained from by permuting the last indices (i.e., those in ).
Concerning case (B), suppose there was a choice of for each , such that (24) was true. Then,
(38) | ||||
so . A similar calculation with integrating out up to indices of shows that .
Now, since is generated by elements of type (A) and (B), we conclude that always implies and , so the definitions of and do not depend on the choice of the coset representative .
∎
Lemma 5.2 (Extended CCR).
Proof.
This follows by a direct calculation using basis coefficients, as in the case .
∎
Lemma 5.3 (Bogoliubov Transformations Conserve Extended CCR).
Proof.
Lemma 5.2 ensures that the CCR hold for . Lemma 37 renders well–definedness of for . The recovery of the CCR for using the Bogoliubov relations (17) works as in the case of being Fock space operators. For instance, we have777It is easy to verify that the first two expressions are well–defined operators . The other expressions are elements of , which can be interpreted as multiplication operators , since the –product of two –elements is again in , see Remark 2.
(41) | ||||
The other two identities in (40) are obtained analogously.
∎
As in the case [19], the definition of our extended implementer is based on a vector , called Bogoliubov vacuum. We first give a definition of for a given and then construct further below. Recall the definition of (31).
Definition 5.1.
Given a Bogoliubov–transformed vacuum state , we define the linear extended Bogoliubov implementer by
(42) |
with and .
5.2 Construction of the Bogoliubov vacuum
The only remaining step for finishing the construction of is to provide a reasonable Bogoliubov vacuum , which we do in Definition 50. By “reasonable”, we mean that is annihilated by all –operators, which we prove in Lemma 51. This property will play an important role in the proof of our main theorem.
Let us quickly explain the heuristics for the choice of . In case , the Bogoliubov vacuum is well–known [19, (61)]: Up to normalization, it consists of an exponential over two–particle wave functions of the kind
(43) |
where indexes a simultaneous eigenbasis of both888We have , so . Further, for , both operators are Hilbert–Schmidt, which allows for a simultaneous eigenbasis. and with eigenvalues (with respect to ), and where and . The orthonormal basis is then given by . Using that is a spectral multiplication by , it is now an easy task to verify that
(44) |
So when comparing with (14), we see that is the integral kernel of an operator which turns into .
When generalizing to the case with arbitrary spectrum , an eigenbasis of will generally no longer exist. However, we still have a projection–valued measure (PVM) with
(45) |
And we are still able to define an operator , which turns into and has an integral kernel : Consider the polar decompositions with respect to ,
(46) |
following from , where functions of are to be understood as spectral multiplications and are anti–unitary operators. Then, is defined as
(47) |
which is clearly bounded by .
Lemma 5.4 (Bosonic Pairs).
For any , we have
(48) |
Further, allows for an integral kernel , so
(49) |
Proof.
Definition 5.2 (Bosonic Bogoliubov Vacuum).
We define by
(50) |
for and for .
In what follows, only the sector will be relevant, so we will drop the index . Further, under the embedding described in Remark 1.
Lemma 5.5 ( Annihilates Bosonic ).
For , we have
(51) |
Proof.
Recall (16) and the definition of the antilinear conjugation operator , which imply
(52) |
The creation term has only contributions of odd sector numbers that evaluate to
(53) |
The annihilation term also has only contributions of odd , which can be evaluated in mode–configuration space. In the following, let and be permutations of and , let , and denote by the sums over all or permutations. Then,
(54) | ||||
In order to evaluate the term in brackets, we use that is the integral kernel of , so
(55) |
For evaluating the other term, we use that the integral kernel of is , so
(56) |
The Bogoliubov relations (17) and polar decompositions (46) now imply
(57) | ||||
with being a densely defined spectral multiplication operator. Now, if is some densely defined operator on and is its adjoint, then for and , we have
(58) | ||||
So
(59) | ||||
Plugging (55) and (59) into (54), we obtain
(60) |
which exactly cancels (53) and establishes the lemma. ∎
5.3 Conditions for Extended Implementability
Before proceeding to the final proof of Theorem 4.1, we first set up some simple conditions for when is an extended implementer. These conditions are analogous to the ones given in [13, Lemma 5.2].
Lemma 5.6 (Conditions for an Extended Implementer).
Let such that . Further, let as in Definition 5.1 be injective (so exists). Then, implements in the extended sense.
Proof.
In order to establish the extended implementation (32), it suffices to show
(61) |
for and
(62) |
where and are arbitrary. In case , we directly compute
(63) |
In case , we obtain
(64) | ||||
On the other hand, conservation of the CCR implies that , so
(65) | ||||
Expressions (64) and (65) agree, which renders the desired equality.
∎
5.4 Proof of Theorem 4.1
Proof of Theorem 4.1.
We have to show that has an implementer , i.e., (32) holds. The operator and the vector are given in Definitions 5.1 and 50, respectively. By Lemma 5.6, implements if and is injective. The property readily follows from Lemma 51. So it remains to establish injectivity of in order to finish the proof.
Since is spanned by vectors of the type , it suffices to show that the set
(66) |
is linearly independent. To do so, we investigate the set step by step.
Let us start by evaluating
(67) |
For even sectors, we have . For evaluating the odd sectors, we use that for , the same arguments as in (54) through (60) yield . Thus, together with the definition of (50), we get
(68) |
Now, implies
(69) |
By Lemma 49, we have , so . Thus, the operator is injective, as is (since ), and linear independence of implies linear independence of .
Now, let us turn to the evaluation of a general . For ,
(70) | ||||
By a similar expansion in terms of commutators, one easily sees that
(71) |
where (“lower–order terms”) is a sum of expressions with and .
Now, linear independence of will follow if we can prove linear independence of the set of “leading–order terms”
(72) |
To see this implication of linear independence, suppose, would be linearly dependent, so there was a linear combination
(73) |
with . By we denote the highest number of consecutively applied creation operators. Then, amounts to a finite linear combination of elements of the kind , where the contribution of terms with is
(74) |
with the sum being nonempty. As , linear independence of would now imply that whenever , which contradicts our premise . So linear independence of implies linear independence of .
Finally, we establish linear independence of by a contradiction. Suppose there was a linear combination
(75) |
with and being the least number of consecutively applied creation operators. Then,
(76) |
i.e., the lowest sectors are unoccupied. Thus, the –sector of our linear combination amounts to
(77) |
Since is linearly independent in , also the set
(78) |
is linearly independent, which implies for all with , establishes the desired contradiction and finishes the proof.
∎
Acknowledgments.
This paper originated from discussions with Andreas Deuchert at the INdAM Quantum Meetings 2022, which were supported by the Istituto Nazionale di Alta Matematica ”F. Severi”. The author was further financially supported by the Basque Government through the BERC 2018-2021 program, by the Ministry of Science, Innovation and Universities: BCAM Severo Ochoa accreditation SEV-2017-0718, as well as by the European Research Council (ERC) through the Starting Grant FermiMath, Grant
Agreement No. 101040991.
References
- [1] R. Haag: The Mathematical Structure of the Bardeen-Cooper-Schrieffer Model Nuovo Cimento 25.2: 287–299 (1962)
- [2] V. Bach, E.H. Lieb, J.P. Solovej: Generalized Hartree-Fock Theory and the Hubbard Model J. Stat. Phys. 76: 3–89 (1994) https://arxiv.org/abs/cond-mat/9312044
- [3] N. Benedikter, P.T. Nam, M. Porta, B. Schlein, R. Seiringer: Optimal Upper Bound for the Correlation Energy of a Fermi Gas in the Mean–Field Regime Commun. Math. Phys. 374: 2097–-2150 (2020) https://arxiv.org/abs/1809.01902v3
- [4] M. Falconi, E.L. Giacomelli, C. Hainzl, M. Porta: The Dilute Fermi Gas via Bogoliubov Theory Ann. Henri Poincaré 22: 2283–2353 (2021) https://arxiv.org/abs/2006.00491v1
- [5] A. Adhikari, C. Brennecke, B. Schlein: Bose-Einstein Condensation Beyond the Gross-Pitaevskii Regime Ann. Henri Poincaré 22: 1163–1233 (2021) https://arxiv.org/abs/2002.03406v4
- [6] F.A. Berezin: The Method of Second Quantization Academic Press (1992)
- [7] P. Marecki: Quantum Electrodynamics on Background External Fields PhD Thesis, University of Hamburg (2003) https://arxiv.org/pdf/hep-th/0312304.pdf
- [8] J. Derezińki: Bosonic quadratic Hamiltonians J. Math. Phys. 58: 121101 (2017) https://arxiv.org/abs/1608.03289v2
- [9] C.G. Torre, M. Varadarajan: Functional Evolution of Free Quantum Fields Classical and Quantum Gravity 16: 2651–2668 (1999) https://arxiv.org/abs/hep-th/9811222
- [10] S. Hollands, R.M. Wald: Local Wick Polynomials and Time Ordered Products of Quantum Fields in Curved Spacetime Commun. Math. Phys. 223: 289–326 (2001) https://arxiv.org/abs/gr-qc/0103074v2
- [11] D. Shale: Linear Symmetries of Free Boson Fields Trans. Amer. Math. Soc. 103: 149–167 (1962)
- [12] D. Shale, W.F. Stinespring: Spinor Representations of Infinite Orthogonal Groups J. Appl. Math. Mech. 14.2: 315–322 (1965)
- [13] S. Lill: Implementing Bogoliubov Transformations Beyond the Shale–Stinespring Condition Preprint (2022) https://arxiv.org/abs/2204.13407
- [14] J.v. Neumann: On infinite tensor products Comp. Math. 6: 1–77 (1939)
- [15] S. Lill: Extended State Space for describing renormalized Fock spaces in QFT Preprint (2020) https://arxiv.org/abs/2012.12608
- [16] Yu.M. Berezansky, Y.G. Kondratiev: Spectral Methods in Infinite-Dimensional Analysis Springer (1995)
- [17] H. Araki: On the Diagonalization of a Bilinear Hamiltonian by a Bogoliubov Transformation Publ. Res. Inst. Math. Sci. 4.2: 387–412 (1968)
- [18] P.T. Nam, M. Napiórkowsi, J.P. Solovej: Diagonalization of bosonic quadratic Hamiltonians by Bogoliubov transformations J. Funct. Anal. 270.11: 4340–4368 (2016) https://arxiv.org/abs/1508.07321v2
- [19] J.P. Solovej: Many Body Quantum Mechanics lecture notes http://web.math.ku.dk/~solovej/MANYBODY/mbnotes-ptn-5-3-14.pdf (2014), visited on Jul 1, 2021