The relation between the symplectic group Sp(4, R) and its Lie algebra: its application in polymer

Guillermo Chac´on-Acosta∗ Departamento de Matem´aticas Aplicadas y Sistemas, Universidad Aut´onomaMetropolitana Cuajimalpa, Vasco de Quiroga 4871, Ciudad de M´exico 05348, MEXICO

Angel Garcia-Chung† Departamento de F´ısica, Universidad Aut´onomaMetropolitana - Iztapalapa, San Rafael Atlixco 186, Ciudad de M´exico 09340, M´exico and Universidad Panamericana, Tecoyotitla 366. Col. Ex Hacienda Guadalupe Chimalistac, C.P. 01050 Ciudad de M´exico, M´exico

In this paper, we show the relation between sp(4, R), the Lie algebra of the symplectic group, and the elements of Sp(4, R). We use this result to obtain some special cases of symplectic matrices relevant to the study of squeezed states. In this regard, we provide some applications in quantum mechanics and analyze the squeezed polymer states obtained from the polymer representation of the symplectic group. Remarkably, the polymer’s dispersions are the same as those obtained for the squeezed states in the usual representation.

Contents

I. Introduction 2

II. Sp(4, R) group analysis and sp(4, R) Lie algebra 4 arXiv:2102.12049v1 [math-ph] 24 Feb 2021 III. Quantum relations and examples 9

A. Relation between sp(4, R) and P(2, R) 9 B. Examples 10 1. Case a, c 6= 0 and b = 0. 10

2. Case c = a = diag(a11, a22) and b =diag(b11, b22). 11 3. Case a = c = 0 and b 6= 0. 12

∗Electronic address: [email protected] †Electronic address: [email protected] 2

IV. Quantum representation and its applications 12

A. Schr¨odingerrepresentation of Sp(2n, R) 13 B. Schr¨odingerrepresentation of the squeeze for a bi-partite system. 14 C. Covariance matrix for squeezed states 17

V. Squeezed states in polymer quantum mechanics 20

VI. Conclusions 24

VII. Acknowledgments 25

VIII. Appendix: calculation of m2n 25

IX. Appendix: Series analysis 27

X. Appendix: Covariance matrix coefficients 29

References 29

I. INTRODUCTION

Squeezed states are broadly used in many areas of physics [1–5]. Of particular interest is the use of these squeezed states in cosmology [6–14], specifically when arguing for the emergence of semi-classical behavior in the early universe. Loop Quantum Cosmology (LQC) [15–19] is another scenario in which squeezed states are relevant. There, squeezed states for a single-mode show some of the features of the quantum bounce and closely approximate solutions to the classical Einstein equations [20–25]. The squeezed states used in LQC are constructed by hand, imposing the Gaussian form of the states to obtain the squeezing nature of the dispersion relations. Moreover, the states describe systems with only one degree of freedom, i.e., single-mode squeezed states [26, 27]. In , squeezed states can be used to improve the sensitivity of measurement devices beyond the usual quantum noise limits [2–5, 28–31]. They are defined by the squeeze operators’ action on coherent states, or the vacuum state [2–4]. These operators are defined within the Fock representation using the annihilation and creation operators, or in the Wigner representation, using the Wigner functional. A particular squeezed state used in quantum optics is the two-mode squeezed state which plays a prominent role in the study of entanglement for 3 bipartite systems. Particularly in the limit when the amount of squeezing is infinitely large, the states become EPR-like states [32]. Based on the relevance that squeezed states play in cosmology, LQC, and quantum optics, one might ask whether there is a relation between them and whether it is possible to obtain squeezed states in LQC the same way squeezed states are defined in quantum optics. Recall that the construction used in LQC for the squeezed states is somewhat artificial and does not correspond to any mechanism in the cosmological events. Hence, exploring whether LQC formulation admits an operator similar to the squeeze operator and whose action on some state yields a squeezed state might pave the way to construct such a mechanism in LQC. To do so, one must consider that in the LQC, the representation of the operators is not weakly continuous, hence the Fock representation is not suitable for the physical description. Instead, the Schr¨odingerrepresentation, which is the scheme inherited from the quantization procedure, seems to be the natural scheme to be considered [33–38]. Despite the Schr¨odinger representation of the squeeze operator might be obtained using the representation of the infinitesimal squeeze operator via the exponential map [39], in LQC this cannot be done. In addition to the mathematical challenge that this operation requires in standard quantum mechanics, this is not possible in LQC because there is no infinitesimal representation of the squeeze operator. Therefore, in the present work, we will use the representation of the symplectic group Sp(2n, R) in polymer quantum mechanics [38]. It is worth mentioning that Polymer Quantum Mechanics (PQM) can be considered as a “toy model” of LQC because they share Hilbert spaces with the same mathematical structures. As a result, the representation of the symplectic group in PQM is mathematically the same as in LQC. To analyze the squeeze operator corresponding to the bipartite squeezed states in LQC, we will provide the relation between the Lie algebra of the symplectic group sp(4, R) and the Lie group Sp(4, R). As far as the authors’ knowledge, this relation has not been reported before. With this result, we show some specific cases and then move to the analysis within polymer quantum mechanics. Also, this relation allows us to describe the single-mode squeeze operator (specifically, the product of two single-mode operators) as a particular case of a symplectic matrix in Sp(4, R). We will show that the squeezed states derived in this way for LQC share the same features as those used in quantum optics. In particular, the correlations’ structure is the same for both the single-mode and the two-mode squeezed states. However, there is no need for a Gaussian-like structure for the initial states upon which the polymer squeeze operators act and such structure is absent in the polymer squeezed states. 4

This paper is organized as follows: in section (II) we calculate the relation between sp(4, R) and Sp(4, R). In section (III) we discuss the isomorphism between sp(4, R) and the second-order polynomial operators P(2, R) and provide some examples. In Section (IV), we show some of the applications of the results given in section III; in particular, we determine the covariance matrix for the squeezed states in standard quantum mechanics. In Section (V), we analyze the squeeze operators’ representation in polymer quantum mechanics and construct the polymer squeezed state. We also calculate the dispersion relations and show that they are equal to those obtained for the standard squeezed states. We give our conclusion in Section (VI).

II. Sp(4, R) GROUP ANALYSIS AND sp(4, R) LIE ALGEBRA

In this section we will detail the relation between an arbitrary element of sp(4, R) and its corresponding element in the group Sp(4, R). This relation is the main result of this section and has not been reported as far as we know. First, let us introduce some preliminary concepts and notation, which we will use throughout the paper, to make the presentation self-contained. 2n Let us begin by considering the Poisson manifold (R , {, }) with Poisson bracket for the coor- dinates qj and momenta pj (j = 1, 2, 3, . . . , n) given by

{qj, qk} = 0, {pj, pk} = 0, {qj, pk} = δjk. (1)

T These coordinates are collected using the array Y~ = (q1, p1, q2, p2, . . . , qn, pn) for which the Poisson bracket (1) takes the form   J 0 ··· 0    0 J ··· 0  n~ ~ T o   Y, Y =   = 1n×n ⊗ J, (2)  . . . .   . . . .    0 0 ··· J where the 0 is the 2 × 2 null matrix, 1n×n is the identity matrix and the matrix J is given by   0 1 J =   . (3) −1 0

2n The group action over the manifold R is

2n 2n   0T T Sp(2n, R) × R → R ; M, Y~ 7→ Y~ = M Y~ , (4) provided that the matrix M satisfies the condition

T (1n×n ⊗ J) = M (1n×n ⊗ J) M , (5) 5

T where M is the transpose matrix. That is, the symplectic group Sp(2n, R) can be defined as the set of 2n × 2n real matrices satisfying (5) and, additionally, its group action on the Poisson 2n 2n manifold (R , {, }) is given by (4). Note that a “coordinatization” of (R , {, }) different from Y~ yields a condition for the symplectic group matrices different to that in (5). To show this, consider T T T T T now the array X~ = (~q , ~p ) where ~q = (q1, q2, . . . , qn) and ~p = (p1, p2, . . . , pn) are the 2n coordinates on the space R . The Poisson bracket for this array is given by   0 1 n ~ ~ T o n×n X, X =   = J ⊗ 1n×n. (6) −1n×n 0

2n 2n   0 0 The group action is now given by Sp(2n, R) × R → R ; Mf, X~ 7→ X~ where X~ is

X~ 0T = Mf X~ T , (7) and the matrix Mf satisfies

T (J ⊗ 1n×n) = Mf (J ⊗ 1n×n) Mf . (8)

Hence, both conditions (5) and (8), can be considered as definitions for the symplectic group in 2n different “coordinatizations” of the phase space R . Naturally, both group actions Mf and M are related via the similarity transformation Γ(n) [4] as

Mf = Γ(n) MΓ−1(n), (9) where Γ(n) is given by

X~ T = Γ(n) Y~ T , (10) and is such that ΓT (n) = Γ−1(n). Since the present work concerns the case where n = 2, it is worth showing the explicit form of Γ(2) which is   1 0 0 0      0 0 1 0  Γ(2) =   . (11)    0 1 0 0    0 0 0 1

2n Having provided the two group actions over the manifold R using different “coordinatizations” and their relation for arbitrary n, let us now focus on the symplectic group Sp(4, R). According to (5) this group is given by 4 × 4 real matrices M for which the following condition holds     J 0 J 0 T   = M   M , (12) 0 J 0 J 6

The matrix M can be written in block form as   AB M :=   , (13) CD where the 2 × 2 block matrices A, B, C and D satisfy the conditions

J = AJAT + BJBT = CJCT + DJDT , 0 = AJCT + BJDT , (14) which result from (12).

The Lie algebra of Sp(4, R), denoted as sp(4, R), is given by 4 × 4 matrices m such that the exponential map [39] of the Lie algebra element m yields symplectic matrices M close to the identity, i.e., 1 1 M = em := 1 + m + m2 + ··· + mn + ... (15) 2 n!

It can be shown that the matrices in sp(4, R) can be written as the product   J 0 m =   L, (16) 0 J where L is a real symmetric matrix written in block form as   a b L =   , (17) bT c and where b is a 2 × 2 real matrix, whereas a and c are also real but 2 × 2 symmetric matrices. If a matrix M can be written as in (15), then its inverse M−1, its transpose MT and the n-power matrix (M)n, can be written respectively as follows     −1 J 0 M = exp −   L, (18) 0 J     T J 0 −1 J 0 M = −   M   , (19) 0 J 0 J    n J 0 (M) = exp   (nL). (20) 0 J

Thus, the Lie algebra multiplication in sp(4, R) is given by the matrix commutator [, ]m. When this multiplication acts on two arbitrary elements m1 and m2 gives the element m3 defined as        J 0 J 0 J 0 m3 := [m1, m2]m =   L1,   L2 =   L3, (21) 0 J 0 J 0 J 7

where the matrix L3 is also a real symmetric matrix with components of the form   T T a1Ja2 + b1Jb2 − a2Ja1 − b2Jb1 a1Jb2 + b1Jc2 − a2Jb1 − b2Jc1 L3 =   , (22) T T T T T T b1 Ja2 + c1Jb2 − b2 Ja1 − c2Jb1 c1Jc2 + b1 Jb2 − c2Jc1 − b2 Jb1 hence, m3 is clearly an element in sp(4, R). Up to this point, we introduced the main concepts and notations required to derive the relation between sp(4, R) and its corresponding Lie group Sp(4, R). Let us proceed then to obtain the explicit relation between the block matrices A, B, C and D and the Lie algebra element L. It is worth noting that the following procedure can be applied to higher-order symplectic groups

Sp(2n, R) for n ≥ 3, being this the main reason for its exposition in this section. Let us collect the even and odd terms of the expansion in (15) as follows

 1 1   1 1  M = 1 + m2 + ··· + m2n + ... + m 1 + m2 + ··· + m2n + ... , (23) 2! (2n)! 3! (2n + 1)! where m2 takes the form   2 −(det a + det b)12×2 Jd m =   , (24) T −Jd −(det b + det c)12×2 and the matrix d is defined as

d = aJb + bJc. (25)

As can be seen from the expansion (23), to obtain the expression for M we need first to determine m2n. In Appendix VIII we obtain the expression for m2n given in Eq. (111). Let us replace this result in the series expansion (23), which, after collecting the even and odd terms, gives the following

A = α(e) + (α(o) − β(o) det b) J a + β(o)J b J c J bT , (26)

B = (γ(o) − β(o) det a) J b + β(e)(J a J b + J b J c) + β(o)J a J b J c, (27)

C = (α(o) − β(o) det c) J bT + β(e)(J bT J a + J c J bT ) + β(o)J c J bT J a, (28)

D = γ(e) + (γ(o) − β(o) det b) J c + β(o)J bT J a J b. (29)

The coefficients α(e), α(o), β(e), β(o), γ(e) and γ(o) were defined in the appendix IX. These expressions for the matrices A, B, C and D link the components of the Lie algebra element L with the corresponding symplectic matrix M and constitute the main result of this section. Note also the “non-linear matrix relation” between the Lie algebra elements and the group elements, particularly the role of the block matrix b. 8

A remarkable and direct application of this result is that it allows us to compute the symplectic eigenvalues of the matrix M. To do so, recall that the characteristic polynomial for a 4 × 4 matrix M with det(M) = 1 can be written in terms of the trace of its first three powers by the expression

(Tr(M))2 − (Tr(M2))  (Tr(M))3 (Tr(M))(Tr(M2)) Λ4 − (Tr(M))Λ3 + Λ2 + − + + 2 6 2 (Tr(M3)) − Λ + 1 = 0. (30) 3 where Λ are the eigenvalues of the arbitrary matrix M and Tr(M) is the trace of the matrix. Using the relations (26)–(29) we obtain that the trace Tr(M) is given by

(e) (e) p  p  Tr(M) = 2(α + γ ) = cosh λ+ + cosh λ− , (31)

(e) (e) where α and γ are given in (129) and (133), respectively, and eigenvalues λ± are given in (120). Moreover, due to the linearity of the trace and the relation (20), we can verify that Tr(Mn) is given by

n  p   p  Tr(M ) = cosh n λ+ + cosh n λ− , (32) from which we obtain the expressions for Tr(M2) and Tr(M3). Inserting all these expressions in (30) for a symplectic matrix of the form (15), and when calculating the roots of the polynomial, we obtain that their eigenvalues are given by

    √ p p − λ+ Λ1 = cosh λ+ − sinh λ+ = e , (33)     √ p p − λ− Λ2 = cosh λ− − sinh λ− = e , (34)     √ p p λ− Λ3 = sinh λ− + cosh λ− = e , (35)     √ p p λ+ Λ4 = sinh λ+ + cosh λ+ = e , (36)

Note that if λ+ > λ−, then the eigenvalues are ordered as Λ1 < Λ2 < Λ3 < Λ4 which coincides with the results of Williamson’s theorem [40, 41]. In [26, 42] an alternative (and different) formulation for each of the symplectic group genera- tors was provided. Our approach, however, not only reproduces to the same expressions for the generators but also provides a direct relation with the Lie algebra matrix generators a, b and c a point that is absent in [26, 42]. As a result, we can relate a broader range of Lie algebra elements with their corresponding group elements. Let us now show some of the relevant matrices and examples in the next section in which this result can be applied. 9

III. QUANTUM RELATIONS AND EXAMPLES

This section provides three examples where the relation between the Lie algebra element and the group element is explicit. However, before proceeding, let us introduce additional concepts and notations (see [4, 39] for more details), which will be relevant for the quantum description.

A. Relation between sp(4, R) and P(2, R)

Consider the Lie algebra formed by second-order (operator) polynomials, denoted by P(2, R).

An arbitrary element sb is given as

i ha11 2 a12 a22 2 sb = − qb1 + (qb1pb1 + pb1qb1) + pb1 + b11qb1qb2 + b12qb1pb2 + b21pb1qb2 + b22pb1pb2+ ~ 2 2 2 c c c i + 11 q2 + 12 (q p + p q ) + 22 p2 . (37) 2 b2 2 b2 b2 b2b2 2 b2

Here, qbj and pbj, with j = 1, 2, are the position and momenta operators satisfying the canonical commutation relations [qbj, pbk] = i~δj,k, and aij, bij and cij are all real coefficients. The reason for this notation is that formally i~sb is a self-adjoint operator to be represented in a Hilbert space H, hence the exponential map esb gives rise to a unitary operator in H. Thus, in this sense, this notation smoothes the way to the quantum representation analysis in section IV.

It is easy to check that sb can be written in the following form   a11 a12 b11 b12     i T  a12 a22 b21 b22  s = − Rb   Rb , (38) b 2   ~  b11 b21 c11 c12    b12 b22 c12 c22 that is a symmetric matrix, and where

  RT = . (39) b qb1 pb1 qb2 pb2

Instead of the matrix commutator, the Lie algebra multiplication in P(2, R) is given by the operator commutator [, ]. Therefore, the Lie algebra multiplication of two elements sb1 and sb2 gives a third element sb3 of the form

i T sb3 = [sb1, sb2] = − Rb L3Rb , (40) 2~ where the matrix L3 is given by (22). Due to L3 is a symmetric matrix, the operator sb3 is clearly in P(2, R). Naturally, this result provides the isomorphism between sp(4, R) and P(2, R), i.e., the 10 map

i T ι : sp(4, R) → P(2, R); m 7→ sb = ι(m) = − Rb L Rb , (41) 2~ and this map preserves the linear properties of both Lie algebras, i.e., it is a Lie algebra isomor- phism.

An implication of this isomorphism is that due to P(2, R) is a Lie algebra isomorphic to sp(4, R), s s then the exponential map of its elements (sb 7→ eb) gives a (quantum) unitary operator (eb) which can be seen as the (quantum) unitary representation of Sp(4, R) as showed in the following diagram

ι sp(4, R) ⇐==⇒ P(2, R)   y y (42) Sp(4, R) −→ Spc(4, R)

We can expect that if a representation of P(2, R) in a Hilbert space H is known, then there is also a representation of Spc(4, R) in H. However, in some scenarios like in polymer quantum mechanics and LQC, it is not possible to obtain the representation of Sb out of the representation of sb in H. The reason is that some elements of P(2, R) cannot be represented in the corresponding Hilbert space. This difficulty can be overcome if we can represent directly the exponential esb instead of its infinitesimal generator sb. This approach was done for the case of polymer quantum mechanics in [38]. Consequently, due to the operators in (37) can be used to describe the dynamics of many physical systems ranging from two decoupled quantum harmonic oscillators to the bipartite squeeze operators, a polymer representation of these operators is possible, as we will show in section V. More details about these aspects will be provided in section V. In the next subsection, we show some of the explicit forms of M.

B. Examples

1. Case a, c 6= 0 and b = 0.

Let us consider the Lie algebra element with b = 0 and a, c 6= 0, which, according to the expression (38), implies that there is no interaction between the subsystems, that is, sb is of the form

i  2 2 2 2 sb = − a11qb1 + a12(qb1pb1 + pb1qb1) + a22pb1 + c11qb2 + c12(qb2pb2 + pb2qb2) + c22pb2 . (43) 2~ 11

In this case, d = 0 and λ+ = − det a and λ− = − det c. After inserting b = 0 and the expressions for λ± in (129)-(134) we obtain the following symplectic matrix √  √ sinh( − det a)  cosh − det a + √ J a 0 M = − det a √ . (44) 1  √ sinh( − det c)  0 cosh − det c + √ J c − det c

As can be seen, both block matrices in (44) are elements of Sp(2, R) hence, the Lie algebra elements given by the parameters a and c can be considered as the Lie algebra generators of Sp(2, R) ⊗

Sp(2, R) ⊂ Sp(4, R). Moreover, the matrix M1 is diagonal if and only if a and c are anti-diagonal matrices, i.e., only when there are no squared terms in (37). An important symplectic matrix of this type is   J 0   , (45) 0 J which is often used to derive the transpose matrix as in (19). One can check that this matrix can π π be obtained from (44) when a = c = diag( 2 , 2 ), i.e.,      J 0 π J 0   = exp   . (46) 0 J 2 0 J

2. Case c = a = diag(a11, a22) and b =diag(b11, b22).

In this case the operator sb is of the form       i 1 2 1 2 1 2 1 2 sb = − a22 pb1 + pb2 + a11 qb1 + qb2 + b11qb1qb2 + b22pb1pb2 , (47) ~ 2 2 2 2 i.e., the sub-systems interact via the matrix b but only with couplings between coordinates qb1qb2 and momenta operators pb1pb2. According to (120), the expression for λ± for this case is

λ± = −(a11 a22 + b11 b22) ± (a11 b22 + a22 b11) = −(a11 ∓ b11)(a22 ∓ b22). (48)

Note that when b11 = ±a11 or b22 = ±a22 the eigenvalues are null and two particular systems emerge with their operators given by     i 1 2 1 2 a11 2 b11 = ±a11 → sb = − a22 pb1 + pb2 + (qb1 ± qb2) + b22pb1pb2 , (49) ~ 2 2 2     i a22 2 1 2 1 2 b22 = ±a22 → sb = − (pb1 + pb2) + a11 qb1 + qb2 + b11qb1qb2 . (50) ~ 2 2 2 Both systems represent two interacting harmonic oscillators with a coupling term in the momenta and the coordinates, respectively. 12

The symplectic matrix, denoted in this case as M2, is given by

 √  √  √  √   cosh λ− +cosh λ+ cosh λ− −cosh λ+ (a22+b22)S−+(a22−b22)S+ (a22+b22)S−+(b22−a22)S+ 2 2 2 2  √  √  √  √    cosh λ− +cosh λ+ cosh λ− −cosh λ+   (b11−a11)S+−(a11+b11)S− (a11−b11)S+−(a11+b11)S−   2 2 2 2  M2 =  √  √  √  √   , cosh λ− −cosh λ+ cosh λ− +cosh λ+  (a22+b22)S−+(b22−a22)S+ (a22+b22)S−+(a22−b22)S+   2 2 2 2   √  √  √  √    cosh λ− −cosh λ+ cosh λ− +cosh λ+  (a11−b11)S+−(a11+b11)S− (b11−a11)S+−(a11+b11)S− 2 2 2 2 (51)

where we introduce the parameters S± as p  sinh λ± S± := p . (52) λ±

3. Case a = c = 0 and b 6= 0.

In this case, the operator sb is of the form i sb = − [b11qb1qb2 + b12qb1pb2 + b21pb1qb2 + b22pb1pb2] , (53) ~ and this system corresponds, as we will see in the next section, to the general case of the squeeze operator for a bi-partite system [4]. Note that in this case, not only the matrices a = c are null, but also the matrix d, which implies that λ+ = λ− = − det b. Once we replace these expressions in (129)-(134) the symplectic matrix takes the form  √ √  cosh − det b sinh√ − det b J b − det b M3 =  √ √  , (54) sinh√ − det b J bT cosh − det b − det b where the block matrices A and D are diagonal matrices. Clearly, when det b < 0 the coefficients of M3 will be given by hyperbolic functions. In case det b > 0 the coefficients are described by trigonometric functions instead. After giving some examples of symplectic matrices obtained through the exponential map, we are ready to show some of their quantum mechanics applications, both in the standard representation and the so-called polymer or loop representation.

IV. QUANTUM REPRESENTATION AND ITS APPLICATIONS

The unitary representation of the group Sp(2n, R) was given by Moshinsky and Quesne in [43]. A review and a historical analysis can be found in [44, 45]. However, to be self-contained, we will 13 show the main aspects of this group’s quantum representation in standard quantum mechanics in the next subsection.

A. Schr¨odingerrepresentation of Sp(2n, R)

The symplectic group is a non-compact group which implies an infinite-dimensional Hilbert 2 n space for its unitary representation. Consider the Hilbert space H = L (R , d~x) where d~x is the standard Lebesgue measure. The unitary representation of Sp(2n, R) is the map

C : Sp(2n, ) → U(H); M 7→ C , (55) b R f bMf

† where Cb is a unitary operator over H, i.e., formally Cb = Cb−1. Note that the group action con- Mf Mf Mf sidered in this map is Mf instead of M, i.e., we used the “coordinatization” given by X~ introduced in section II. Hence, in order to obtain a quantum (unitary) representation of a given symplectic matrix M we first have to transform it into the other group action Mf using Eq. (9) with the corresponding matrix Γ(n) given by (10) or Γ(2) for Sp(4, R) given in (11). The map Cb is given by the integral operator Z C Ψ(~x) = d~x0C (~x,~x0)Ψ(~x0), Ψ(~x) ∈ H, (56) bMf Mf and the kernel C (~x,~x0) of this integral is Mf

i [~xT De Be −1~x−2~x0T Be −1~x+~x0T Be −1Ae ~x0] e 2~ C (~x,~x0) = . (57) Mf q (2πi~)n det Be

According to [43], this representation results from imposing two conditions on the operators C . The first one is given by bMf

 T   T  ~q ~q b −1 −1 b Cb   Cb = Mf   , (58) Mf ~T Mf ~T pb pb and relates the symplectic group elements M with the operators C . Here, ~q := (q , q ,..., q ) f bMf b b1 b2 bn ~ and pb := (pb1, pb2,..., pbn) are the coordinate and momenta operators associated to the Heisenberg Lie algebra of the system. The second condition is that

 † C · C = 1, (59) bMf bMf b where 1 is the identity operator and this results in the unitarity of C . b bMf 14

The factor det B in (57) gives rise to a well define operator even in the case where the matrix B is singular (for more details see [43, 44]). Finally, it is worth mentioning that this representation (56) is valid for the entire symplectic group and not just for those elements close to the group identity. Since the fundamental operators are unbounded the condition (58) only holds in a subspace given by the domain of the operators qbj and pbj in H. To obtain a condition valid in the full Hilbert space, we are forced to introduce the exponentiated version of qbj and pbj, that is to say, the Weyl algebra. Briefly, the Weyl algebra is a C∗-unital algebra whose generators, denoted by Wc(~a,~b), are related with qbj and pbj with the following relation

 T T  i ~a ~q +~b ~p Wc(~a,~b) := e ~ b b , (60) and such that the real arrays ~a = (a1, a2, . . . , an) and ~b = (b1, b2, . . . , bn), which have dimensions

[aj] = momentum and [bj] = position, label the Weyl algebra generators.

The standard Schr¨odingerrepresentation of qbj and pbj is now used to obtain a representation for the generators Wc(~a,~b) in H given by

~ i ~a ~b T i ~a~x T ~ Wc(~a, b)Ψ(~x) = e 2~ e ~ Ψ(~x + b), (61) and such that the canonical commutation relations give rise to the Weyl algebra mutiplication

i ~T ~ T ~ ~ − (~a1 b2 −b1 ~a2 ) ~ ~ Wc(~a1, b1)Wc(~a2, b2) = e 2~ Wc(~a1 + ~a2, b1 + b2). (62)

Combining (58) and (60) to obtain the exponentiated version of (58) yields

C W (~a,~b)(C )−1 = W (~a DT −~b CT , −~a BT +~b AT ), (63) bMf c bMf c e e e e where Ae , Be , Ce and De are the block matrices in Mf. This relation allows us to obtain a representation of the symplectic group in the Hilbert space used in polymer quantum mechanics and in loop quantum cosmology [38]. We are now ready to show, in the next subsections, some of the applications of the representation of Sp(4, R) given by (56) and (57).

B. Schr¨odingerrepresentation of the squeeze operator for a bi-partite system.

The squeeze operator Sb(ζ) for a bi-partite system is given by the exponential map

Sb(ζ) = esbζ , (64) 15 where the operator sbζ , is given by 1   s := ζ∗a a − ζa†a† . (65) bζ 2 b1b2 b1b2

Here, ba1 and ba2 are the annihilation operators for each of the sub-systems, say, 1 and 2, of the † † bi-partite system, ba1 and ba2 are their adjoint operators respectively and ζ is a complex number labelling the amount of squeezing. The operator Sb(ζ), when acting on the vacuum state of the bi-partite quantum harmonic oscillators, gives a family of squeezed states labelled by ζ. The operators in (65) are in the Fock representation, hence, let us transform (65) to the

Schr¨odingerrepresentation described with operators qb1, qb2, pb1 and pb2. The relation between these representations is given by

1 qbj i ljpbj † 1 qbj i ljpbj baj = √ + √ , baj = √ − √ , (66) 2 lj 2 ~ 2 lj 2 ~ q for j = 1, 2 and l := ~ where m and ω stand for the masses and the frequencies of the j mj ωj j j † oscillators. Inserting these expressions for baj and baj in (65) the operator sbζ takes the following form   1 ~ ζy l2 ζx l1 ζx l1l2 ζy sbζ = qb1qb2 − qb1pb2 − pb1qb2 − pb1pb2 , (67) 2i~ l1l2 l1 l2 ~ where ζx and ζy are the real and imaginary parts of ζ. We now rewrite this operator in the form

   ~  i ~ T ~ T 0 b Rb1 sζ = − (Rb1 , Rb2 )     , (68) b 4 T ~ ~ b 0 Rb2 where the matrix b is the following   ~ ζy l2 ζx l l − l b =  1 2 1  . (69) − l1 ζx − l1l2 ζy l2 ~

Using the isomorphism ι−1 defined in (41) we obtain that the corresponding Lie algebra element −1 mζ = ι (sbζ ) is given by     J 0 0 b mζ =     . (70) 0 J bT 0

Note that the Lie algebra matrix mζ isomorphic to the squeeze operator sbζ , is of the type given in the third case (III B 3). 16

To obtain the symplectic matrix associated to this Lie algebra element, we insert mζ and its expressions for a, b and c in (54). This results in the following symplectic matrix   l1 l1l2 cosh(r) 0 − sinh(r) cos(φ) l − sinh(r) sin(φ)  2 ~   0 cosh(r) − sinh(r) sin(φ) ~ sinh(r) cos(φ) l2   l1l2 l1  Ms(r, φ) =   ,  l2 l1l2   − sinh(r) cos(φ) l − sinh(r) sin(φ) cosh(r) 0   1 ~  − sinh(r) sin(φ) ~ sinh(r) cos(φ) l1 0 cosh(r) l1l2 l2 (71) iφ where r and φ are defined as ζ = re . Matrix Ms(r, φ) can be considered as the classical symplectic 2 2 2 transformation such that when represented in L (R , d ~x), gives rise to the quantum operator Sb(ζ). Naturally, this means also that the unitary representation of Sb(ζ) in the Schr¨odingerrepresentation

sbζ is given by CbMs , i.e., CbMs = e .

It is worth to mention that although the expression (71) depends on the proper lengths l1 and l2, the matrix Ms(r, φ) is ~−independent, i.e., it is entirely a classical object. Also, matrix Ms produces classical squeezing but of course, adapted to the classical phase space, which in this 4 case is (R , {, }). To illustrate the squeezing and the rotation properties of the matrix Ms as a canonical transformation for different values of r and φ we consider its action on a circular trajectory

(q1(t), p1(t), q2(t), p2(t)) where, qj(t) = cos(t) qj + sin(t) pj and pj(t) = − sin(t) qj + cos(t) pj, for j = 1, 2.

p1

p1

2

2

1

1

q1 -1.5 -1.0 -0.5 0.5 1.0 1.5 q1 -2 -1 1 2

-1 -1

-2 -2

(a) Squeezing a circular trajectory. (b) Rotating a squeezed trajectory with r = 0.6.

FIG. 1: In both figures, the solid, the dashed and the dotted lines correspond to: (a) r = 0, r = 0.3 and r = 0.6,

respectively and (b) to φ = 0, φ = π/4 and φ = π/2, respectively. 17

The action of Ms on the trajectories is explicitly of the form

 0    q1(t) q1(t)      0     p1(t)   p1(t)    = Ms (r, φ)   . (72)  0     q2(t)   q2(t)      0 p2(t) p2(t)

0 0 In figure 1 we showed the plot of (q1(t), p1(t)). As expected, we note in Fig. (1a), that the amount of squeezing r squeezes the circular trajectory. Recall that symplectic transformations also preserve the area, hence the trajectories are squeezed but the area is preserved. On the other hand, the rotation angle φ, as showed in (1b), rotates the trajectories and also preserves the area.

Finally, observe that Ms is given in the Y~ “coordinatization” which is not suitable for its quantum representation. To make it suitable, let us provide the expression for the matrix Mfs, which is given by

 l1 sinh(r) cos(2φ) l1l2 sinh(r) sin(2φ)  cosh(r) − l 0 −  2 ~   − l2 sinh(r) cos(2φ) cosh(r) − l1l2 sinh(r) sin(2φ) 0   l1 ~  Mfs =   . (73)  ~ sinh(r) sin(2φ) l2 sinh(r) cos(2φ)   0 − l l cosh(r) l   1 2 1  − ~ sinh(r) sin(2φ) 0 l1 sinh(r) cos(2φ) cosh(r) l1l2 l2 This expression will be used to explore the analog of the bipartite squeeze operator in polymer quantum mechanics in section V.

C. Covariance matrix for squeezed states

Now we will show the relation between the covariance matrix, denoted by V(2), and the sym- plectic matrix M. Let us consider the state |Ψ i ∈ L2( n, d~x) related with the symplectic matrix f Mf R Mf as

|Ψ i = C |0i, (74) Mf bMf R where |0i = d~x Ψ0(~x) |~xi is the state |0i = |0i1 ⊗|0i2 ... |0in, and the ket |0ij is the vacuum state of the j-th quantum harmonic oscillator. Note that this construction can be extended to other 2 n states in L (R , d~x) and not only for |0i. However, for simplicity in our exposition, let us consider the simplest example of the covariance matrix for C |0i. bMf To obtain the covariance matrix we first calculate the following amplitude

† hΨ |Wc(~a,~b)|Ψ i = h0|Cb Wc(~a,~b) Cb |0i, (75) Mf Mf Mf Mf 18 where Wc(~a,~b) is the Weyl-algebra generator introduced in (60). Combining (56), (61) and the Gaussian form of the vacuum state of the system given by n-decoupled harmonic oscillators, we obtain the following expression for the amplitude in (75)     1  T ~a  ~ ~ hΨM|Wc(~a, b)|ΨMi = exp − ~a b Λ   , (76) f f  4 ~b  where the matrix Λ is given by   1 2 2 L 0 T Λ := Mf  ~  Mf , (77) 0 L−2 and L = diag(l1, l2, . . . , ln), where lj was defined earlier (66). The covariance matrix V(2) has components given by   hΨ |x x |Ψ i 1 hΨ | {x , p } |Ψ i (2) M bj bk M 2 M bj bk M V =  f f f f  , (78) 1 hΨ | {p , x } |Ψ i hΨ |p p |Ψ i 2 Mf bj bk Mf Mf bj bk Mf and these components can be obtained from (76) using the following relations

hΨ |x x |Ψ i = − 2∂2 hΨ |W (~a,~b)|Ψ i| , (79) Mf bj bk Mf ~ aj ak Mf c Mf ~a,~b=0 1 2 2 ~ hΨ | {xj, pk} |Ψ i = − ∂ hΨ |Wc(~a, b)|Ψ i| ~ , (80) 2 Mf b b Mf ~ aj bk Mf Mf ~a,b=0 1 2 2 ~ hΨ | {pj, xk} |Ψ i = − ∂ hΨ |Wc(~a, b)|Ψ i| ~ , (81) 2 Mf b b Mf ~ bj ak Mf Mf ~a,b=0 2 2 hΨ |pj p |Ψ i = − ∂ hΨ |W (~a,~b)|Ψ i| . (82) Mf b bk Mf ~ bj bk Mf c Mf ~a,~b=0

Remarkably, the resulting expression for V(2) in terms of the symplectic matrix Mf is   2 (2) 1 L 0 T V = Mf   Mf , (83) 2 2 −2 0 ~ L and this shows the direct relation between the covariance matrix V(2) for the state |Ψ i and the Mf symplectic matrix M associated with the unitary operator C . Moreover, if we now consider the f bMf definition (8), it can be shown that 2 V(2) is actually a symplectic matrix. Let us apply this formula ~ to some of the systems considered before.

Consider the matrix M1 given in (44). Using (9) we obtain the expression for Mf1 which then is replaced in (83) giving rise to the following covariance matrix

 (2) (2)  V11 0 V13 0    (2) (2)  (2) 1  0 V22 0 V24  V =   . (84) 1 2  (2) (2)   V13 0 V33 0   (2) (2)  0 V24 0 V44 19

Its components are given in the appendix (X) and in the particular case where a11 = a22 = c11 = (2) c22 = 0, the covariance matrix V1 reduces to   2 2a12 l1e 0 0 0    0 l2e2c12 0 0  (2) 1  2  V1 =  2  . (85) 2  ~ −2a12  0 0 2 e 0  l1   2  ~ −2c12 0 0 0 2 e l2

We use this result to derive the uncertainties in the coordinates ∆xj for j = 1, 2 which according to (85) are given by

q αj 2 2 lje ∆xj := hΨ |x |Ψ i − hΨ |xj|Ψ i = √ , (86) Mf bj Mf Mf b Mf 2 where due to the symmetry of the vacuum wavefunction we have hΨ |x |Ψ i = 0. This can Mf bj Mf be verified calculating the first derivatives in (76). Here, for simplicity we make α1 = a12 and

α2 = c12. Note the remarkably property of the squeezed states like in (74) which is that ∆xj can be smaller than the proper length of the vacuum state lj when αj < 0.

Similarly, the uncertainties in the momenta ∆pj, are q ∆p := hΨ |p2|Ψ i − hΨ |p |Ψ i2 = √ ~ , (87) j Mf bj Mf Mf bj Mf α 2lje j

~ which can also be smaller than when αj > 0 and also, hΨ |pj|Ψ i = 0 as the previous case. lj Mf b Mf Nevertheless, both uncertainties satisfy Heisenberg’s uncertainty principle: ! l eαj  ∆x ∆p = j√ √ ~ = /2. (88) j j α ~ 2 2lje j Another interesting covariance matrix is the one related with the bipartite squeeze operator (73) derived in the previous subsection. Inserting (73) in (83) yields

 l2 cosh(2r)  1 − l1l2 sinh(2r) cos(2φ) 0 − l1~ sinh(2r) sin(2φ) 2 2 2l2  l2 cosh(2r)   − l1l2 sinh(2r) cos(2φ) 2 − l2~ sinh(2r) sin(2φ) 0  (2)  2 2 2l1  V (r, φ) =  2 2  ,  l2~ sinh(2r) sin(2φ) ~ cosh(2r) ~ sinh(2r) cos(2φ)  0 − 2  2l1 2l1 2l1l2   2 2  l1~ sinh(2r) sin(2φ) ~ sinh(2r) cos(2φ) ~ cosh(2r) − 0 2 2l2 2l1l2 2l2 (89) and this allows us to determine the correlation between the second moments of the subsystem 1 and the subsystem 2 e2r e−2r hΨ | (x ± x )2 |Ψ i = l2 + l2 ∓ 2l l cos(2φ) + l2 + l2 ± 2l l cos(2φ) , (90) Mf b1 b2 Mf 4 1 2 1 2 4 1 2 1 2 2e2r 2e−2r hΨ | (p ± p )2 |Ψ i = ~ l2 + l2 ± 2l l cos(2φ) + ~ l2 + l2 ∓ 2l l cos(2φ) .(91) Mf b1 b2 Mf 2 2 1 2 1 2 2 2 1 2 1 2 4l1 l2 4l1 l2 20

In the particular case where l = l = l and φ = π , the uncertainties (∆x ) and (∆x ) for 1 2 2 1 Ψe 2 Ψe the state Ψe are correlated as follows

2 2 2 (∆x1) + (∆x2) = l cosh(2r). (92) Ψe Ψe

These are the main results, at the standard quantum mechanics level, which we want to show regarding the representation of the symplectic group in quantum mechanics. There are others ap- plications like the analysis of the Bohmian trajectories for bipartite squeezed states, the analysis of the fidelity for bipartite or tripartite squeezed states, and others which are currently in preparation. For now, let us move to the analysis of the squeezed states in polymer quantum mechanics given in the next section.

V. SQUEEZED STATES IN POLYMER QUANTUM MECHANICS

Polymer quantum mechanics [33, 34, 36, 37], is a quantization scheme which can be considered as a “toy model” looming from loop quantum cosmology. Hence, exploring the nature and properties of squeezed states in polymer quantum mechanics will help study those scenarios in loop quantum cosmology where such states might play a significant role. For this example, we will consider a system with two degrees of freedom and both will be polymer quantized. Therefore, the Hilbert space of the entire system is given by

(1) (2) Hpoly = Hpoly × Hpoly, (93)

(j) where the Hilbert spaces Hpoly with j = 1, 2 are of the form

(j) 2 (j) Hpoly = L (R, dpBohr), (94)

(j) where R is the Bohr compactification of real line and dpBohr is the Bohr measure (see [36] for more details). This Hilbert space resembles the momentum representation used in the standard quantum mechanics.

An arbitrary state in this Hilbert space Hpoly is given by

X i ~xT ~p ~ j Ψ(p1, p2) = Ψ~xj e , (95) {~xj }

(1) (2) j=n where {~xj} is a shorthand notation for the graph {(xj , xj )}j=1 associated with the state Ψ(~p).

In this notation, the array ~p = (p1, p2) denotes the momentum variables for the system 1 and 2, 21

respectively. The coefficients Ψ~xj provide the value for the norm of the state which is given by

X 2 ||Ψ(~p)|| = |Ψ~xj | , (96) {~xj } hence these coefficients are different from zero and the sum converges (they are non-null over countable number of points in the graph {~xj}). This norm arises from the inner product

Z L1 Z L2 1 ∗ hΨ|Φi = lim Ψ (~p) Φ(~p)dp1dp2, (97) L ,L →∞ 1 2 4 L1L2 −L1 −L2 which for the specific case of the plane waves takes the form of the Kronecker delta

i ~xT ~p i ~x0T ~p he ~ |e ~ i = δ~x,~x0 . (98)

This inner product is the main signature of the polymer quantization as it violates the Stone- von Neumann theorem. Consequently, polymer quantum mechanics is not unitarily equivalent to the standard Schr¨odingerrepresentation. Moreover, in polymer quantum mechanics there is no hence infinitesimal spatial translations cannot be implemented. Nevertheless, we can obtain a representation for the , which in the present case is given by

∂ X (1) i ~xT ~p q Ψ(p , p ) = i Ψ(p , p ) = − Ψ x e ~ j , (99) b1 1 2 ~ 1 2 ~xj j ∂p1 {~xj }

∂ X (2) i ~xT ~p q Ψ(p , p ) = i Ψ(p , p ) = − Ψ x e ~ j . (100) b2 1 2 ~ 1 2 ~xj j ∂p2 {~xj }

Despite these peculiarities with the non-regularity of the polymer representation, the represen- tation of the symplectic group Sp(2n, R) on the Hilbert space of polymer quantum mechanics was provided recently by one of the authors in Ref. [38]. There, the representation is given by the map (poly) (poly) (poly) Cb : Sp(2n, R) → L(Hpoly), Mf 7→ Cb , where the linear operator Cb acts on Hpoly as Mf Mf

Z L1 Z L2 (poly) 1 (poly) Cb Ψ(~p) = lim C (~p,~p0)Ψ(~p0)d~p0. (101) Mf L ,L →∞ Mf 1 2 4 L1 L2 −L1 −L2

The polymer kernel C(poly)(~p,~p0) is given by Mf

(poly) 0 T − 1 − i ~pT Be De −1~p X i ~pT ~x− i ~p0T De T ~x+ i ~xT De Ce T ~x C (~p,~p ) = det(De Ae ) 4 e 2~ e ~ ~ 2~ , (102) Mf ~x

T − 1 and note that when the factor det(De Ae ) 4 6= 1 it implies that this representation is not unitary [38]. Recall that one of the main features of loop quantum cosmology is its intrinsic length scale given by the Planck length. As a toy model, polymer quantum mechanics does not have an intrinsic 22 length scale. However, it admits a length scale that mimics some of the features of loop quantum cosmology. This length scale is introduced at hand and is called polymer scale, usually denoted by µ. This polymer scale constitutes the analog of minimum length for polymer quantum mechanics models, and therefore, it can be considered as a lower bound for the uncertainties. In the present analysis, each system admits a polymer scale µ1 and µ2 when the dynamics is considered. Let us now consider the following questions: (1) is it possible to have polymer states such that their uncertainties are lower than the polymer scale? Furthermore, (2) do the correlations found in (90) have an analog in polymer quantum mechanics?

To answer these questions let us consider the matrix Mf1 with a11 = a22 = c11 = c22 = 0, used (2) to calculate the covariance matrix V1 in (85), but now with a12 = −r1 and c12 = −r2. The explicit form is Mf1   e−r1 0 0 0    −r2   0 e 0 0  Mf1 =   . (103)  r1   0 0 e 0    0 0 0 er2

(poly) The action of the group element Cb on an arbitrary polymer state (95) gives the following Mf1 state

i  −r (1) −r (2) (poly) X e 1 p1x +e 2 p2x ~ j j Ψe (p1, p2) = Cb Ψ(p1, p2) = Ψ~xj e . (104) Mf1 Mf1 {~xj }

Using the representation of the position operators qb1 and qb2 given in (99) and (100) we obtain the dispersion relations

(∆x ) = e−r1 (∆x ) , (∆x ) = e−r2 (∆x ) , (105) 1 Ψe 1 Ψ 2 Ψe 2 Ψ which show that the squeezed polymer state is indeed squeezed by a factor e−r1 or e−r2 . Con- sequently, if the initial dispersion of the polymer state is given by (∆x1)Ψ or (∆x2)Ψ, then the (poly) squeeze operator Cb gives rise to a polymer state Ψe (104) whose dispersion is smaller than that Mf1 of the initial polymer state Ψ. Moreover, due to there is no upper bound for the parameter r, these dispersion relations can be smaller than the corresponding polymer scales µ1 and µ2. Let us now consider the analog of the correlations (92) but for polymer states. To do so, (poly) consider the polymer representation Cb of the symplectic matrix (73) corresponding to a bi- Mfs (poly) partite system. The action of Cb on an arbitrary polymer state (95) is given by Mfs

i h (1) (2)  (2) (1) i (poly) X cosh(r) x +sinh(r) x p1+ cosh(r) x +sinh(r) x p2 ~ j j j j Ψe s = Cb Ψ(p1, p2) = Ψ ~xj e , (106) Mfs {~xj } 23 where again the representation in (101) was used. We combine this result with the representation of the position operators in (99) and (100) and obtain the following relations

2 2 2 2 (∆x1) − (∆x2) = (∆x1)Ψ − (∆x2)Ψ, ∀ Ψ ∈ Hpoly, (107) Ψe s Ψe s

π where the conditions l1 = l2 and φ = {0, 2 , π} were imposed. Remarkably, this result not only is independent of the parameter r (which labels the amount of squeezing) but also applies to any polymer state Ψ ∈ Hpoly. As can be seen, the difference of dispersions squared is conserved, regardless of the amount of squeezing. Also, note that l1 and l2 are considered group parameters and have no relation to the dynamics, i.e., we are considering general states in Hpoly. The same applies for φ. Let us now consider pure and symmetric polymeric states. The pure states are those that can be written as the following product     i (1) i (2) (p) X (1) p1x X (2) p2x Ψ (p1, p2) =  Ψ e ~ j   Ψ e ~ j  . (108)  x(1)   x(2)  (1) j (2) j {xj } {xj }

(1) (2) (s) Secondly, both lattices {xj } and {xj } are symmetric, i.e., for every positive point 0 < xj ∈ (s) (s) (s) (s) (s) {xj } there exist a negative point 0 > xj0 ∈ {xj }, such that xj + xj0 = 0, and the states are also symmetric which implies that Ψ(s) = Ψ(s) . These states are the analog of the states x(s) x(s) j j0 described with even functions in the standard quantum mechanics. The dispersion relation for squeezed pure symmetric polymer states is given by

2 2 2 2  (∆x1) + (∆x2) = cosh(2r) (∆x1) (p) + (∆x2) (p) , (109) Ψe s Ψe s Ψ Ψ which takes the form

2 2 2 (∆x1) + (∆x2) = l cosh(2r), (110) Ψe s Ψe s √ when the dispersion of the pure states are ((∆x1)Ψ(p) = (∆x2)Ψ(p) ) = l/ 2. Notably, Eq. (110) is the same as that obtained in (92) for the Schr¨odingerrepresentation. This shows that the correlations present in the standard quantum mechanics using the C operator for both symplectic bMf matrices Mf1 and Mfs are the same to those obtained in polymer quantum mechanics using the (poly) operator Cb . Mf 24

VI. CONCLUSIONS

In this paper we provided the direct relation between the Lie algebra sp(4, R) and the symplectic group Sp(4, R). The expression shows the link between the block matrices A, B, C and D with those of the Lie algebra a, b and c given in the Eqs. (26)-(29). This result has not been reported before and applies to the full Lie algebra sp(4, R) of the symplectic group Sp(4, R). Such relation allows us to obtain some important symplectic matrices that were used in subse- quent sections. In the first example for a, c 6= 0 and b = 0, we show that the corresponding symplec- 0 0 0 0 tic matrix M1, given in (44), can be written as (M1 ⊗ 1) · (1 ⊗ M2), where 1, M1, M2 ∈ Sp(2, R). 0 0 Here, M1 and M2 are symplectic matrices acting over each of the sub-systems with coordinates

(q1, p1) and (q2, p2) respectively. In the case a = c = diag(a11, a22) and b = diag(b11, b22) the symplectic matrix M2 in (51), describes two coupled harmonic oscillators with interaction terms labeled by the coefficients of the matrix b. Finally, in the equation (54) we showed the symplectic matrix M3 for the case in which a = c = 0 and a general form of the matrix b. In section (IV) we analyzed the classical description of squeeze operators. We showed that the s symplectic matrix Ms is the classical analog of the squeeze operator S(ζ) = ebζ = C . Also, we b bMfs remarked the isomorphism between the Lie algebra sp(4, R) and P(2, R). Additionally, the general form of the covariance matrix V(2) for the squeezed vacuum state |Ψ i was derived using the Mf

Weyl algebra representation and the symplectic matrix M1. The components of this covariant matrix were used to calculate the dispersion relations (86) and (87) for the particular case where a11 = a22 = c11 = c22 = 0. As is already known, these dispersions can be smaller than the vacuum characteristic length for the harmonic oscillators. They also satisfy the Heisenberg uncertainty principle as was shown in (88). We then calculated the covariance matrix for the symplectic matrix

Ms corresponding to the classical analog of the bipartite squeeze operator. With this matrix we determined the correlation (92). We also provided the general expressions for these correlations in equations (90) and (91). Applying the previous results, it is also possible to represent operators in non-regular Hilbert spaces that are non-unitarily equivalent to the Fock-Schr¨odingerrepresentation, so in section (V) we analyze polymer quantized systems. We calculated the dispersion relation for an arbitrary polymer state using the representation of the symplectic group in polymer quantum mechanics. (poly) We obtained that the polymer representation of the squeeze operator given by Cb , yields a Mf1 dispersion relation Eq. (105), which can be smaller than those of the initial state. This implies (poly) that Cb is indeed a polymer squeeze operator and (104) describes a polymer squeezed state. Mf1 25

(poly) On the other hand, the polymer representation of the bipartite squeeze operator given by Cb Mfs was used to derive the polymer correlations (107) and (110). The first correlation (107) shows that the difference of the dispersions’ square is preserved and is independent of the initial polymer state. (poly) Clearly, this result only holds for Cb so a symplectic matrix different than the one used in (103) Mfs will produce a different result. In the case of (110), the result has the same form as the standard (poly) correlation in Eq. (92), hence, the polymer representation of Cb can be used to construct Mfs correlated squeezed states for bipartite polymer systems. Naturally, this brings some questions like whether there is any mechanism in nature, say loop quantum cosmology or the interior of a black hole, from which a polymer squeezed state can be created. Moreover, establishing that squeezing is a property also present in non-regular representations questions its role in the classicality of some cosmological models. As we showed, it is possible (poly) to construct entangled polymer states using Cb . Such polymer entangled states correlations Mfs satisfy a relation identical to the one obtained in the standard quantum mechanics. In this case, the states are polymer bipartite squeezed states, similar to those used in the quantum description of the inflaton field. Finally, it is worth to mention that the polymer squeezed states obtained as a result of the representation of Sp(4, R) in Hpoly, given in (104, 106), differ from those reported in the LQC literature [20–25]. In these references, a Gaussian form of the states is considered, whereas in our case, the polymer state (104) is a general polymer state. In LQC, the squeezed states are constructed by hand due after imposing some conditions to achieve the squeezed nature of the dispersion relations. In our case, the squeezed state results from the action of the squeeze operator. The results of this work open the doors for studying the entangled states of matter and geometry and the role that their correlations might play in some physical scenarios.

VII. ACKNOWLEDGMENTS

I thank Academia de Matem´aticasand Colegio de F´ısica,UP, for the support and enthusiasm.

VIII. APPENDIX: CALCULATION OF m2n

In this appendix the expression for the matrix m2n is obtained. To do so, recall that the matrix m2 is formed by four 2 × 2 block matrices where the upper left and the lower right matrices are multiples of the identity matrix 12×2. The upper right block is the matrix Jd whereas the lower 26 left is −JdT . Notably, we found that this block structure is preserved after exponentiating the matrix m2 an integer number of times. That is to say, the n-power of matrix m2 yields a new matrix (m2)n given by   2 n αn12×2 βnJd (m ) =   . (111) T −βnJd γn12×2

It is this pattern the one to be considered when this procedure is applied to higher order symplectic groups.

The coefficients αn, βn and γn, are to be determined and depend on the values of the matrices a, b, c and d. For n = 1, these coefficients are given by the factors in the block matrices of m2 given in (24) and can be directly defined as

α1 := −(det a + det b), β1 := +1, γ1 := −(det c + det b). (112)

To calculate αn, βn and γn for arbitrary n, first note that they can be generated with the (n − 1)-power of the matrix T as     αn α1       n−1    βn  = T  β1  , (113)     γn γ1 where the matrix T is given by   α1 β1 det d 0     T =  β1 γ1 0  . (114)   0 β1 det d γ1

The calculation shows that Tn−1 is a matrix of the form   n−1 ~T n−1 U 0 T =   , (115) T n−2 Pn−2 −j j n−1 ~u γ1 j=0 γ1 U γ1 where ~0 = (0, 0) and ~u = (0, β1 det d) and matrix U is given by   α1 β1 det d U =   . (116) β1 γ1

Then, using (113) we have the following relation for the coefficients

    n−2   αn n−1 α1 n T n−2 X −j j α1   = U   , γn = γ1 + ~u γ1 γ1 U   . (117) βn β1 j=0 β1 27

In order to calculate Un−1 we need to diagonalize the matrix U. Let P be the matrix diago- nalizing U, then

−1 U = PD0 P , (118) where the matrix P is   (λ+−γ1) (λ−−γ1) β k1 β k2 P =  1 1  . (119) k1 k2

The real arbitrary parameters k1 and k2 result from the diagonalization procedure. Its values will be automatically cancelled as part of the calculation of Un−1 further below. The eigenvalues of U, denoted by λ±, have the following expression

α + γ 1q λ = 1 1 ± (α − γ )2 + 4β2 det d, ± 2 2 1 1 1 det a + det c + 2 det b 1p = − ± (det a − det c)2 + 4 det d. (120) 2 2 and the diagonal matrix D0 is   λ+ 0 D0 =   . (121) 0 λ−

We now take the n − 1 power of U given in (118) to obtain the following result      −1 (λ+−γ1) (λ−−γ1) n−1 (λ+−γ1) (λ−−γ1) n−1 β k1 β k2 λ+ 0 β k1 β k2 U =  1 1     1 1  , (122) n−1 k1 k2 0 λ− k1 k2 which, when combined with the result in (117) together with the expression for ~u, gives

n n (λ+ − det b − det c)λ+ − (λ− − det b − det c)λ− αn = , (123) p 2 2 (det a − det c) + 4β1 det d n n λ+ − λ− βn = , (124) p 2 2 (det a − det c) + 4β1 det d  n n  (λ+ − det b − det c)λ− − (λ− − det b − det c)λ+ γn = . (125) p 2 2 (det a − det c) + 4β1 det d These are the final expressions for the coefficients in (m2)n.

IX. APPENDIX: SERIES ANALYSIS

In this appendix we calculate the series expansion terms. To do so, note that once the expression for (m2)n is inserted the expansion (23) and the even and odd terms are collected, the matrix 28

M(a, b, c) takes the form     (e) (e) (o) (o) α 12×2 β Jd α 12×2 β Jd M(a, b, c) =   + m   , (126) (e) T (e) (o) T (o) −β Jd γ 12×2 −β Jd γ 12×2 where the following coefficients are given by

+∞ +∞ +∞ X 1 X 1 X 1 α(e) := 1 + α , β(e) := β , γ(e) := 1 + γ , (127) (2n)! n (2n)! n (2n)! n n=1 n=1 n=1

+∞ +∞ +∞ X 1 X 1 X 1 α(o) := 1 + α , β(o) := β , γ(o) := 1 + γ . (128) (2n + 1)! n (2n + 1)! n (2n + 1)! n n=1 n=1 n=1 We now insert (123), (124) and (125) in the relations (127) - (128) to obtain

h p  p i (det c − det a) cosh λ+ − cosh λ− (e) 1 h p  p i α = cosh λ+ + cosh λ− + , (129) 2 2p(det a − det c)2 + 4 det d

 p  p   p  p  1 sinh λ+ sinh λ− (det c − det a) sinh λ+ sinh λ− α(o) = + + − ,  p p  p 2  p p  2 λ+ λ− 2 (det a − det c) + 4 det d λ+ λ− (130)

(e) 1 h p  p i β = cosh λ+ − cosh λ− , (131) p(det a − det c)2 + 4 det d

 p  p  1 sinh λ+ sinh λ− β(o) = − , (132) p 2  p p  (det a − det c) + 4 det d λ+ λ−

h p  p i (det c − det a) cosh λ+ − cosh λ− (e) 1 h p  p i γ = cosh λ+ + cosh λ− − , (133) 2 2p(det a − det c)2 + 4 det d

 p  p   p  p  1 sinh λ+ sinh λ− (det c − det a) sinh λ+ sinh λ− γ(o) = + − − ,  p p  p 2  p p  2 λ+ λ− 2 (det a − det c) + 4 det d λ+ λ− (134)

where we have to recall the expression for the eigenvalues λ± in (120). 29

X. APPENDIX: COVARIANCE MATRIX COEFFICIENTS

In this appendix we show the explicit form of the coefficients of the covariance matrix V(2). √ √ 2 2 4 2 2 2 (2) sinh ( − det a) a12l1 + a22~ a12l sinh(2 − det a) √ V = − + 1 √ + l2 cosh2( − det a), (135) 11 l2 det a − det a 1 √ 1 √ 2 2 4 2 2 2 (2) sinh ( − det c) c12l2 + c22~ c12l sinh(2 − det c) √ V = − + 2 √ + l2 cosh2( − det c), (136) 22 l2 det c − det c 2 √ 2 √ 2 4 2 4 2 (2) sinh(2 − det a) a22~ − a11l1 a12 sinh ( − det a) a11l1 + a22~ V = √ + , (137) 13 2l2 − det a l2 det a √ 1 √ 1 2 4 2 4 2 (2) sinh(2 − det c) c22~ − c11l2 c12 sinh ( − det c) c11l2 + c22~ V = √ + , (138) 24 2 2 2l − det c l2 det c √ 2 √ √ 2 2 4 2 2 2 2 2 (2) sinh ( − det a) a11l1 + a12~ a12 sinh(2 − det a) cosh ( − det a) V = − + ~ √ + ~ (139), 33 2 2 2 l1 det a l − det a l1 √ 1 √ √ 2 2 4 2 2 2 2 2 (2) sinh ( − det c) c11l2 + c12~ c12 sinh(2 − det c) cosh ( − det c) V = − + ~ √ + ~ .(140) 44 2 2 2 l2 det c l2 − det c l2

[1] Walls, D. F., “Squeezed states of light”, Nature, 306, 5939, 1983, Springer. [2] Braunstein, S. L. and Van Loock, P., “Quantum information with continuous variables”, Reviews of Modern Physics, 77, 2, 2005. [3] Walls, D. F. and Milburn, G. J., “Quantum optics”, 2007, Springer Science & Business Media. [4] Adesso, G., Ragy, S. and Lee, A. R., “Continuous variable quantum information: Gaussian states and beyond”, Open Systems & Information Dynamics, 21, 2014, World Scientific. [5] Schnabel, R., “Squeezed states of light and their applications in laser interferometers”, Physics Reports, 684, 2017, Elsevier. [6] Grishchuk, L. P. and Sidorov, Y. V., “Squeezed quantum states of relic gravitons and primordial density fluctuations”, Physical Review D,(42), 1990. [7] Polarski, D. and Starobinsky, A. A., “Semiclassicality and decoherence of cosmological perturbations”, Classical and Quantum Gravity,(13), 1996. [8] Lesgourgues, J., Polarski, D. and Starobinsky, A. A., “Quantum-to-classical transition of cosmological perturbations for non-vacuum initial states”, Nuclear Physics B, 497, 1997. [9] Kiefer, C., Polarski, D. and Starobinsky, A. A., “Quantum-to-classical transition for fluctuations in the early universe”, International Journal of Modern Physics D, 7, 1998. [10] Guth, A. H. and Pi, S., “Quantum mechanics of the scalar field in the new inflationary universe”, Physical Review D, 32, 1985. [11] Martin, J. and Vennin, V., “Quantum discord of cosmic inflation: Can we show that CMB anisotropies are of quantum-mechanical origin?”, Physical Review D, 93, 2016. 30

[12] Albrecht, A., Ferreira, P., Joyce, M. and Prokopec, T., “Inflation and squeezed quantum states”, Physical Review D, 50, 1994. [13] Gasperini, M. and Giovannini, M., “Quantum squeezing and cosmological entropy production”, Clas- sical and Quantum Gravity, 10, 9, 1993. [14] Grain, J. and Vennin, V., “Squeezing formalism and canonical transformations in cosmology”, arXiv preprint arXiv:1910.01916, 2019. [15] Ashtekar, A., Bojowald, M. and Lewandowski, J., “Mathematical structure of loop quantum cosmol- ogy”, Advances in Theoretical and Mathematical Physics, 7, 2, 2003. [16] Bojowald, M., “Canonical gravity and applications: cosmology, black holes, and quantum gravity”, Cambridge University Press, 2010. [17] Bojowald, M., “Quantum cosmology: a fundamental description of the universe”, vol. 835, Springer Science & Business Media, 2011. [18] Ashtekar, A., Pawlowski, T. and Singh, P., “Quantum nature of the big bang: an analytical and numerical investigation”, Physical Review D, 73, 12, 2006. [19] Ashtekar, A., Pawlowski, T. and Singh, P., “Quantum nature of the big bang: Improved dynamics”, Physical Review D, 74, 8, 2006. [20] Mielczarek, J. and Piechocki, W., “Gaussian state for the bouncing quantum cosmology”, Physical Review D, 86, 8, 2012. [21] Gazeau, J. P., Mielczarek, J. and Piechocki, W., “Quantum states of the bouncing universe”, Physical Review D, 87, 12, 2013. [22] Diener, P., Gupt, B., Megevand, M. and Singh, P., “Numerical evolution of squeezed and non-Gaussian states in loop quantum cosmology”, Classical and Quantum Gravity, 31, 16, 2014. [23] Diener, P., Gupt, B. and Singh, P., “Numerical simulations of a loop quantum cosmos: robustness of the quantum bounce and the validity of effective dynamics”, Classical and Quantum Gravity, 31, 2014. [24] Taveras, V., “Corrections to the Friedmann equations from loop quantum gravity for a universe with a free scalar field”, Physical Review D, 78, 2008. [25] Corichi, A. and Montoya, E., “Coherent semiclassical states for loop quantum cosmology”, Physical Review D, 84, 2011. [26] Milburn, G. J., “Multimode minimum uncertainty squeezed states”, Journal of Physics A: Mathematical and General, 17, 4, 1984. [27] Gerry, Ch., Knight, P. and Knight, P. L., Introductory quantum optics, (2005), Cambridge university press. [28] Simon, R. Sudarshan, E. C. G. and Mukunda, N., “Gaussian pure states in quantum mechanics and the symplectic group”, Physical Review A, 37, 8, 1988. [29] Ma, X. and Rhodes, W., “Multimode squeeze operators and squeezed states”, Physical Review A, 41, 9, 1990. [30] Pirandola, S., Serafini, A. and Lloyd, S., “Correlation matrices of two-mode bosonic systems”, Physical 31

Review A, 79, 2009. [31] Weedbrook, C. et. al., “Gaussian quantum information”, Reviews of Modern Physics, 84, 2, 2012. [32] Einstein, A., Podolsky, B. and Rosen, N. “Can quantum mechanical description of physical reality be considered complete?” Phys. Rev. 47, (1935). [33] Ashtekar, A., Fairhurst, S. and Willis, J. L., “Quantum gravity, shadow states and quantum mechanics”, Classical and Quantum Gravity, 20, 6, 2003. [34] Corichi, A., Vukasinac, T. and Zapata, Jos´eA., “Polymer quantum mechanics and its continuum limit”, Phys. Review D, 76, 4, 2007. [35] Flores-Gonz´alez,E. and Morales-T´ecotl,H. A. and Reyes, J. D., “Propagators in polymer quantum mechanics”, Annals of Physics, 336, 2013. [36] Velhinho, J. M., “The quantum configuration space of loop quantum cosmology”, Classical and Quan- tum Gravity, 24, 14, 2007. [37] Pawlowski, T., Villase˜nor,E. J. S. and Barbero, J. F., “Separable Hilbert space for loop quantization”, Phys. Review D, 90, 6, 2014. [38] Garcia-Chung, A.“The symplectic group in Polymer Quantum Mechanics,” Physical Review D, 101, 10, 2020. arXiv:2003.00388 [gr-qc]. [39] Hall, M., “The theory of groups”, 2018, Courier Dover Publications [40] Arnold, V. I, Mathematical methods of classical mechanics, Springer, 1989. [41] De Gosson, M. A. Symplectic geomtry and quantum mechanics, Birkhauser, 2006 [42] Casta˜nos,O. and Moshinsky, M. Matrix representation of the generators of symplectic algebras. I. The case of sp (4, R), Journal of Physics A: Mathematical and General, 20, 3, 1987. [43] Moshinsky, M. and Quesne, C., “Linear canonical transformations and their unitary representations”, Journal of Mathematical Physics, 12, 8, 1971. [44] Wolf, K., “Development of linear canonical transforms: a historical sketch”, in Linear Canonical Trans- forms, 3-28, 2016, Springer. [45] Torre, A., “Linear ray and wave optics in phase space: bridging ray and wave optics via the Wigner phase-space picture”, 2005, Elsevier.