Open access peer-reviewed chapter

Perspective Chapter: Squeezing and Entanglement of Two-Modes Quantum X Waves

Written By

Ali Saif M. Hassan, Waleed S.A. Hasan and Mohamed A. Shukri

Submitted: 17 June 2023 Reviewed: 02 July 2023 Published: 24 August 2023

DOI: 10.5772/intechopen.1002263

From the Edited Volume

Quantum Entanglement in High Energy Physics

Oliver K. Baker

Chapter metrics overview

39 Chapter Downloads

View Full Metrics

Abstract

Quantum theory of generalized X waves with orbital angular momentum in dispersive media, and the interaction of quantized X waves in quadratic nonlinear media were studied in (J. opt,20,065201 (2018)). We present a kind of phase matching, which is called velocity phase matching, and this phase matching can be used for determining the length of the nonlinear crystal or the interaction time in the experiment setup, to produce X waves with particular velocity v. Moreover, we introduce more analysis for the dependence of squeezing of X waves on its spectral order, and for spectral orders j>0, we predict the existence of a characteristic axicon aperture for maximal squeezing. Then, we find the quantum squeezed state of the down-converted state generated by the χ2-nonlinear process. Finally, we detect their entanglement using a criterion of separability.

Keywords

  • entanglement
  • X waves
  • squeezed state
  • nonlinear quantum optics
  • phase matching
  • scalar field

1. Introduction

Based on the work of Tesla, German physicist Konstantin Meyl created a new unified field and particle theory. In Meyl’s theory [1], vortices are used to explain quantum and classical physics, mass, gravitation, the constant speed of light, neutrinos, waves, and particles. With the help of this model, the properties of subatomic particles may be precisely estimated. The unified equation can also be used to derive well-known equations. He offers resources for carrying out an experiment by Tesla that proves the existence of scalar waves. Simple energy vortices in the form of particles are what scalar waves are. The electromagnetic, eddy current, potential vortex, and special distributions are all covered by the unified field theory. According to Meyl, the field—which creates particles through decay or conversion—always comes first. Energy particles (i.e., potential vortices) were not included in the theory because they are not recognized by classical physics. Quantum physics attempted to explain everything in terms of vortices which is why it is incomplete. Gravitation is from the speed of light difference caused by proximity that, proportional to field strength, decreases the distance of everything for the field strength [2]. The spatially localized vortex structures include pseudo-nondiffracting (P-N) vortex beams and their superpositions and provide additional effects interesting for theoretical research, experiments, and applications [3]. As we see, this kind of vortex in its quantum form can connect to Meyl’s vortex to explain unified field and particle theory of Meyl. Also, it can connect to the localizability of Busch [4] and others for elementary particle physics to solve the theory of measurement in relativistic quantum mechanics problems as stated in Ref. [5]. Vortex solutions produced in some particle physics models like extended Abelian Higgs model [6], Galaxies and their dark-matter halos [7].

The primary characteristic that unifies numerous advancements in quantum physics is entanglement [8]. Entanglement is required, as a prime example, to distinguish between classical and quantum physics using Bell inequalities [9, 10]. Another way to think of entanglement is as the tool that permits real quantum protocols like teleportation [11] and Bell inequalities-based cryptography [12]. When a quantum method, like Shor’s algorithm, yields a meaningful performance advantage over a classical computer, large entanglement is anticipated to be present in quantum registers [13]. It is obvious that entanglement is essential to comprehending and using quantum physics. Therefore, it makes sense to examine the creation of entanglement at its most fundamental source, which are the particle physics theories of fundamental interactions. A Bell inequality would never be violated in nature if the quantum theory of electromagnetic, or QED, never produced electron entanglement. This suggests that at a fundamental level, quantum unitary development must produce entanglement. The gauge principle, which states that the physical laws are invariant despite internal local rotations for particular symmetry groups, has become widely accepted as a way to explain the fundamental interactions seen in nature. With regard to electroweak and strong interactions, the Standard Model uses a Lagrangian that is essentially constrained by gauge symmetry requirements, except from quantum gravity. It is only logical to continue looking for a principle that is even more basic. Previous research has looked at entanglement’s function in particle physics. Orthopositronium has been demonstrated in Ref. [14] to decay into 3-photon states that can be used in Bell-like experiments that reject classical physics more quickly than the conventional 2-particle Bell inequality. Bell inequalities have also been considered in regard to neutrino oscillations [15], kaon physics [16, 17], and their relationship with the characterization of T-symmetry violation [18]. The S-matrix formalism has been used to examine how entanglement varies in an elastic scattering process in Ref. [19]. A recent study on entanglement in deep elastic scattering is also worth mentioning [20]. Bell inequalities have also been discussed in regard to quantum correlations in CMB radiation [21].

Squeezed states, and two-mode-squeezed states, are foundational for describing the basis nature of quantum mechanical phenomena. Their quantum entanglement plays a central role in quantum physics, experiments of quantum optics, and quantum information science [22, 23, 24, 25, 26, 27]. Nonlinear quantum optics, in particular, parametric down-conversion (PDC), gives us ways to study issues in the foundations of quantum mechanics that are not readily available using other technologies [27, 28]. The quantum communication encode information into the polarization degrees of freedom of photons [12] or the orbital angular momentum (OAM) [29, 30, 31, 32, 33, 34, 35]. The important problem related to multi-modes quantum communications is the diffraction and dispersion of the electromagnetic waves [30]. Electromagnetic waves packet is usually subjected to diffraction and dispersion. Diffraction create a broadening in space during propagation of the wave and dispersion create a broadening in the time during the propagation. Ultimately, these effects are connected with the bounded nature of the wave spectrum and, therefore, to its finite energy content [36]. A great effort has been done in studying the effect of atmosphere turbulence in free-space communication [31, 37, 38, 39, 40]. Maxwell’s equations admit diffraction-free and dispersion-free solutions, the so-called localized waves [3]. In particular, such solutions in the monochromatic domain are the Bessel beams [41], and in the pulsed domain, the most renowned localized waves are the X waves, first introduced in acoustics in 1992 by Lu and GreenLeef [42]. Despite the great work in the literature concerning X waves, the investigations of their quantum properties are very few [43, 44]. Generalization of the traditional X waves to the case of OAM-carrying X wave and the coupling between angular momentum and the temporal degrees of freedom of ultrashort pulses have been investigated [45]. Very recently, quantum and squeezing of X waves with OAM in nonlinear dispersive media have been proposed [46, 47], and they can open a new direction for free-space quantum communication and different areas of physics. In this work, we state the propagation of a scalar electromagnetic field in a linear dispersive medium. We present a kind of phase matching called velocity phase matching and introduce a relation between the interaction time and the velocity of X wave as well as a relation between the length of the χ2-nonlinear crystal and the velocity of X wave generated by the spontaneous parametric down-conversion (SPDC) process in Section 2. We study the SPDC process in a quadratic medium, in particular the dependence of squeezing of the down-converted state generated by the χ2-nonlinear crystal on the spectral order of quantum X waves in Section 3. We introduce the quantum squeezed form of the state generated by the SPDC process and its entanglement in Section 4. Finally, the results are summarized in Section 5.

Advertisement

2. Propagation of a generalized X wave in dispersive media and its quantization

We consider the propagation of a scalar electromagnetic field in a linear dispersive medium with refractive index n=nω. Applying the paraxial and slowly varying envelope approximation to the envelope function Art of an electric field

Ert=2ε0n2Arteikzwt,E1

where Art varies slowly with zand k=nωω/c, which are the propagation coordinate and the wave number, respectively.

The field envelope Art satisfies the following eq. [48],

iAt+iωAzω22Az2+ω2kΔ2A=0,E2

where ω=c2dk/ and ω=c2d2k/dω2 are the first and the second order dispersion, respectively.

The solution of Eq. (2) can be written as a superposition of the generalized X waves as follows [47]:

Art=m,pdvCmpveiv22ωtψm,pvRζ,E3

where ψm,pvRζ is the generalized X wave of OAM number m, spectral order p and velocity v defined as,

ψm,pvRζ=0fpαJmωk/ωαReiαv/ωζeimθ,E4

with R=x2+y2, ζ=zω+vt which is the co-moving reference frame associated to the X waves, and fpα=k/π2ω1+pΔαLp12αΔeαΔ which is the spectrum function, where Lp1x is the generalized Laguerre polynomials of the first kind of order p [48], and Δ is the reference length related to the spatial extension of the spectrum [47]. The quantized field of Eq. (3) can be written as [47],

Ârt=m,pdveiMv22tωm,pvψm,pvRζâm,pv+h.c.,E5

where h.c. denotes to the Hermitian conjugate, ωm,pv=v2/2ω and M=/ω is the mass of X wave. The Hamiltonian operator can be written as follows [47]:

Ĥ=m,pdvωm,pvâm.pvâm,pv+12,E6

with the following usual bosonic commutation relations of creation and annihilation operators of the field,

âm,pvân,qu=δm,nδp,qδuv,âm,pvân,qu=0=âm,pvân,qu.E7

Now, we turn our attention to the quadratic nonlinear process involving X waves, particularly to the spontaneous parametric down-conversion (SPDC) [28]. The interaction Hamiltonian of this process, with a practical case ρ=k1ω2/k2ω11, can be written in two different forms [47]:

The first form is for the time-dependent interaction Hamiltonian which can be written as follows [47]:

I=χA1A2+χA2A1,E8
χA1A2=χd3rÂ1rtÂ2rtE9
χA1A2=χm,p,n,qdudveiu2+v2t2ω2ωm,puωn,qv×d3rψm,puRζ1ψn,qvRζ2âm,pub̂n,qv,E10
ĤIt=m,p,qdudvχm,p,qu+vωm,puωm,pveiFuvtâm,pub̂m,qv+h.c.,E11

with

Fuv=12ω2uv+vuω1ω2,E12

and the interaction function is

χm,p,qu+v=1mχΔ2ω21+p1+qu+vLp1u+vΔ2ωLq1u+vΔ2ωeu+vΔ2ωθu+v,E13

where θu+v, is the Heaviside step function [48]. The second form is for the finite length interaction Hamiltonian which can be given by Ref. [47]

ĤIz=dtχA1A2+χA2A1.E14

The final form of ĤIz can be written as follows:

ĤIz=m,p,qdudvΞm,p,quvωm,puωm,qveiΛuvzâm,pub̂m,qv+h.c.,E15

where Λuv=2uv/ωu+v, and Ξm,p,quv, which is the modified vertex function can be given by Ref. [47],

Ξm,p,quv=1m2χΔ2ω21+p1+qu2+v2u+vLp1u2+v2Δu+vωLq1u2+v2Δu+vωeu2+v2Δu+vω.E16

To calculate the state of the system after the interaction, we apply Schwing-Dyson expansion of the propagator expiĤIt/, and consider only the first term of the expansion [49]. For the time-dependent form of the interaction Hamiltonian stated in Eq. (11), the final state after the χ2-nonlinear interaction can be expressed as follows [47]:

ψ1t=i0tĤIτ0.E17

Substituting Eq. (11) into the above equation and after some calculations, the final state can be written in the following form:

ψ1t=m,p,qdudvGm,p,quvuvtm,p,u;m,q,v,E18

where m,p,u;m,q,vâm,pub̂m,qv0,0, the amplitude Gm,p,quv is defined as

Gm,p,quv=iωm,puωm,qvχm,p,qu+vFuv,E19

and uvt=eiFuvt1. The above state represents the superposition of two particles corresponding to the modes ω1 and ω2 (generated by the nonlinear interaction) traveling with velocities u and v, respectively. For the second form of the interaction Hamiltonian stated in Eq. (15), the final state of the system can be written as

ψ1L=i0LdzĤIz0.E20

Substituting Eq. (15) into the above equation, the state of the field at the output face of the crystal can be written as follows:

ψ1L=m,p,qdudvm,p,quvuvLm,p,u;m,q,v,E21

where uvL=eiΛuvL1, and

m,p,quv=i2ωm,puωm,qvωu+vuvΞm,p,quv.E22

The transition probability for the field to be in such a state after the time-dependent interaction with the χ2-nonlinear crystal is proportional with uvt2 as follows:

Pt(uvt)2,E23

similarly, the transition probability for the finite length interaction of the χ2-nonlinear crystal is proportional with uvL2 as follows [47]:

PL(uvL)2.E24

Explicitly, uvt2 can be expressed as follows:

uvt2=21cosFuvt.E25

From the above equation, we notice that the transition probability reaches its maximum, when uvt2 reaches its maximum, and that occurs when the condition Fuvt=2m+1π, with m=0,1,2,3,, is achieved. The above condition is considered as a new phase matching condition, called velocity matching condition over the velocities of the two-modes X waves generated from SPDC process. This condition requires that

u=2km+vΔω2v+Δω,E26

where Δω=ω1ω2, and km=2m+1πω/t. For the case, when Δω=0, we get uv=km=2m+1πω/t, and the interaction time can be given by t=2m+1πω/uv. When u=v, regardless of that Δω=0, we get v2=km=2m+1πω/t.

Thus, the interaction time can be written in terms of the velocity of X waves modes as follows:

t=2m+1πωv2.E27

For the finite length of the nonlinear crystal, the velocity matching condition is given by Ref. [47],

u=knvvkn,E28

where kn=n+12πω/L, with n=0,1,2,. Therefore, the velocity matching condition fixes either the length of the crystal L or the relative velocity of the two modes X wave involved in the SPDC process as, uv/u+v=kn=n+1/2πω/L, and the length of the crystal L can be given by,

L=n+12πωu+vuv.E29

For u=v, the above equation becomes

L=2n+1πωv.E30

From Eqs. (27) and (30), we note that the ordinary relation between the length and the time is achieved (L=vt). This result can be used for determining the length of the nonlinear crystal in the experimental setup, to produce X waves with velocity v. This is the first result of our work.

Advertisement

3. Squeezing of quantum X waves

Now, we can study the general case of the squeezing effect in the SPDC process, in particular, the dependence of squeezing of the down-converted state, generated by χ2-nonlinear media, on the spectral orders of quantum X waves, for the case, when the velocities of the two-modes X waves are equal u=v, and their spectral orders are equal as well p=q=j. The interaction time Hamiltonian in Eq. (11) becomes

ĤIt=mjdvχmj2vωvâmjvtb̂mjvt+h.c.,E31

where âmjvt=eiFvt/2âmjv, and b̂mjvt=eiFvt/2b̂mjv. Also, the interaction function in Eq. (13) becomes

χmj2v=21mχΔ2vω21+jLj1vΔωLj1vΔωevΔω.E32

And the modified vertex function in Eq. (16) can be rewritten as follows:

Ξmjv=21mχΔ2vω21+jLj1vΔωLj1vΔωevΔω.E33

From Eqs. (32) and (33) we notice that the interaction function and the modified vertex function have the same form. In the interaction picture, we consider the time evolution controlled by ĤIt. Thus, the equations of motion for âmjvt and b̂mjvt are:

dâmjvtdt=1iâmjvtĤIt,E34
db̂mjvtdt=1ib̂mjvtĤIt.E35

Substituting Eq. (31) into the above equations, we obtain

dâmjvtdt=2iωvχmj2vb̂mjvt,E36
db̂mjvtdt=2iωvχmj2vâmjvt.E37

The general solutions of these equations are [50],

âmjvt=coshβmjvtâmjv+sinhβmjvtαmjβmjvb̂mjv,E38
b̂mjvt=coshβmjvtb̂mjv+sinhβmjvtαmjβmjvâmjv.E39

Here, we consider the expression αmj=2iωvχmj2v, with ωv=v2/2ω. To find the squeezing parameter ξmjv=βmjvt, we use the relation βmjv=αmjαmj, then we can write the squeezing parameter as follows:

ξmjv=21mχt1+jΔv˜3ω3Lj1v˜ω2ev˜ω,E40

with v˜=vΔ. The above equation represents the squeezing parameter which depends on the OAM number m, the spectral order j, and the velocity of the two-modes X waves v. Now, we can show the normalized squeezing parameter modulus ξ¯mj=Δξmj4χt as a function of the normalized velocity v˜ω for different values of j,(j=0,1,2,3,4,5), and fixing the time t so that 4χtΔ=1, when m=0, as shown in Figure 1.

Figure 1.

Plot of the normalized squeezing parameter modulus ξ¯mj=Δξmjχ in function of the normalized velocity v˜ω, for different values of the spectral order j where, (a) for j=0, (b) for j=1, (c) for j=2, (d) for j=3, (e) for j=4, (f) for j=5, with fixing the time t, so that 4χtΔ=1, and m=0.

Figure 1 shows the dependence of the squeezing on the spectral order j of X waves. This dependence can be explained in the following points:

  1. The maximum value of the squeezing parameter ξ¯mj and the corresponding normalized velocity v˜ω increases as the spectral order j increases.

  2. The number of the peak values of squeezing depends on the spectral order j and this number can be given by j+1, where the smallest peak is the first and the biggest one is the last.

  3. The squeezing parameter depends on the velocity of X wave as shown in Eq. (40) where the amount of squeezing parameter produced by SPDC process will be maximized at an optimal velocity for every value of the spectral order j, and this velocity can be given by vopt=njω/Δ, where n0=3, n1=6, n2=9.36, n3=12.88, n4=16.49, n5=20.156 for j=0,1,2,3,4,5, respectively as shown in Figure 1. Figure 1(a) shows that for j=0 as in Ref. [46], and the corresponding optimal axicon angles for different values of j can be given by cosθjopt=Δ/njλ. For example, given a nondiffracting pulse with a duration of Δt=8 fs, a carrier wavelength of λ=850 nm and assuming that χ21012 m/V, the optimal axicon angles θjopt that maximize the squeezing for j=0,1,2,3,4,5 will take the following values respectively, θ0opt20, θ1opt62, θ2opt72, θ3opt77, θ4opt80, θ5opt82. Then, we can evaluate the corresponded maximal squeezing parameters to get the following values respectively, (ξm0100s1, ξm1280s1, ξm2520s1, ξm3820s1, ξm41160s1, ξm51550s1).

To illustrate the effect of the OAM on the squeezing more, we can introduce the quadrature operators X̂mjvt=âmjvt+âmjvt, and Ŷmjvt=iâmjvtâmjvt. For e=αmj/βmj=1 or ϕ=0, we get X̂mjvt=eξmjvX̂mjv0, and Ŷmjvt=eξmjvŶmjv0, and the variance of them is given by ΔX̂mjvt=eξmjvŶmjv0. This shows that the SPDC interaction Hamiltonian for the generalized X waves acts as a two-modes squeeze operator which will be illustrated more in the next section. Noticeably, the effect of the OAM number m changes only the sign of the squeezing parameter ξmjv, that is, the squeezed quadrature changes depending on the parity of the orbital angular momentum number m. In particular, if m is even number, ξmjv>0, the squeezing accrues on the Y quadrature. On the other hand, if m is an odd number, ξmjv<0 and the squeezing occurs on X quadrature as in Ref. [46] for j=0, and for j=1,2,3. this effect can be shown as in Figure 2.

Figure 2.

Quadrature space representation of the squeezed down-converted state in the case of even [panel (a)] and odd [panel (b)] values of the OAM number m with a normalized velocity v˜ω=6,9.36,12.88. for a spectral order j=1,2,3, respectively.

Advertisement

4. The two-modes squeezed state of X wave and its entanglement

In analogy to traditional case [51, 52], we look for the eigenvalues of Hamiltonian operator, represented by Eq. (6). Let us introduce the N-particle states m,j,v;N as states with N field excitation with velocity v in the traveling mode (i.e., the X wave) ψm,pvrt with energy Mv2/2 [47]. The eigenstates m,p,v,N of the photon (particle) number operator N̂v=âmpvâmpv are called Fock states, where âmpv is the annihilation operator of the single mode in Eq. (7). These states are orthogonal

nquNmpvNδnmδpqδuv,E41

and complete. The number state m,p,v,N can be generated from the vacuum state (i.e., the eigenstate of N̂v with the eigenvalue equal to zero) as

m,p,v,N=âmpvNN!0.E42

The expectation value of the electric field operator is zero on the particle number operator eigenstates [47],

mpvNÂrtmpvN=0.E43

However, the Fock states m,p,v,N are not normalizable [47], since their representation in configuration state rtmpvN=ψmpvrt carries infinite energy. The particle number states of the two modes AB with particle numbers NANB are defined as

m,j,v;NAA=âmjvNA0ANA!,E44

and

m,j,v,NBB=b̂mjvNB0BNB!E45

Together with the conditions for the vacuums, âmjv0A=0 and b̂mjv0B=0.

The operators âmjv and b̂mjv obey the bosonic commutation relations in Eq. (7). Now, we can define the two-mode squeeze operator, which is a unitary operator, using the normalized squeezing parameter ξ¯mjv=1m21+jv˜3ω3Lj1v˜ω2ev˜ω, as

Vξ¯mj=expeξ¯mjâmjvb̂mjveξ¯mjâmjvb̂mjv.E46

The two-mode squeezed vacuum TMSV state can be obtained from the vacuum state as [52],

0,0,ξ¯mjAB=Vξ¯mj0,0=coshξ¯mj1N=0etanhξ¯mjNm,j,v,N;m,j,v,N.E47

Also, we can define the annihilation operators crossing over the separation of part A and part B as [53],

Âξ¯mj=Vξ¯mjâmjvVξ¯mj=coshξ¯mjâmjv+etanhξ¯mjb̂mjvB̂ξ¯mj=Vξ¯mjb̂mjvVξ¯mj=coshξ¯mjb̂mjv+etanhξ¯mjâmjv.E48

The new field operators satisfy the bosonic commutation relations in Eq. (7)

Âξ¯mjÂξ¯mj=B̂ξ¯mjB̂ξ¯mj=δm,mδj,jδvv,E49

and

Âξ¯mjB̂ξ¯mj=Âξ¯mjB̂ξ¯mj=0.E50

The TMSV state corresponds to the vacuum of the new field operators, as we have used âmjv0=0, b̂mjv0=0, Eq. (47), Eq. (48), and the unitary property of the squeeze operator, we get Âξ¯mj0,0,ξ¯mj=0, and B̂ξ¯mj0,0,ξ¯mj=0.

The two-mode number state is defined by the number state of nonlocal modes as [53],

NA,NB,ξ¯mj=Âξ¯mjNAB̂ξ¯mjNBNA!NB!0,0,ξξ¯mj.E51

In our case, we have NA=NB=1, so we obtain

1,1,ξ¯mj=Âξ¯mjB̂ξ¯mj0,0,ξ¯mj.E52

The state in the above equation represents a two-particles squeezed state of the two modes X wave. The superposition of the two particles corresponding to the two modes ω1 and ω2 see Eq. (18) (when u=v) can be written as

ψ1t=mjdvGmjvvtâmjvb̂mjv0,0,E53

where vt=eiFut1, and Gmjv=iωvχmj2v/Fv, with χmj2v as in Eq. (32). The superposition of the two particles squeezed state corresponds to the state in Eq. (53) is

ψξ¯mj1t=Vξ¯mjψ1t=mjdvGmjvvtVξ¯mjâmjvb̂mjv0,0.E54

Using Eq. (48), the above state can be written in the form

ψξ¯mj1t=mjdvGmjvvtÂξ¯mjB̂ξ¯mj0,0,ξ¯mj,E55

which is a superposition of the squeezed states 1,1,ξ¯mj in Eq. (52).

Now, for calculating the entanglement of the state in Eq. (55), we consider the separable conditions based on the operators that form SU2 algebra and SU11 algebra [53]. The measurements of quadrature moments of the fourth order are sufficient to verify the entanglement of this state. Let us define the operators,

Jz=12NANB,Kx=12âmjvb̂mjv+âmjvb̂mjv,Ky=12iâmjvb̂mjvâmjvb̂mjvE56

where N̂A=âmjvâmjv and N̂B=b̂mjvb̂mjv are number operators of the local modes.

The relations of the annihilation and creation operators âmjv, âmjv, b̂mjv, and b̂mjv with the new field operators in Eq. (48) can be represented by

âmjv=coshξ¯mjÂξ¯mjetanhξ¯mjB̂ξ¯mjâmjv=coshξ¯mjÂξ¯mjetanhξ¯mjB̂ξ¯mjb̂mjv=coshξ¯mjB̂ξ¯mjetanhξ¯mjÂξ¯mjb̂mjv=coshξ¯mjB̂ξ¯mjetanhξ¯mjÂξ¯mjE57

The separable condition is [53],

Δ2Ky14Δ2Jz14Kx2E58

where Ô=trÔρ denotes the expectation value of the observable Ô with respect to state ρ, ΔÔ=ÔÔ, and Δ2Ô=Ô2Ô2. The state we use in Eq. (58) is the superposition of simultaneous eigenstates of the number operators Âξ¯mjÂξ¯mj and B̂ξ¯mjB̂ξ¯mj. From the definition of Jz in Eq. (56), we have

Âξ¯mjÂξ¯mjB̂ξ¯mjB̂ξ¯mj=NANB=2JzE59

This implies that the two-mode squeezed (TMS) number states are eigenstates of Jz and Δ2Jz=0.

Also, we can write

Kx=12cosh2ξ¯mjÂξ¯mjηB̂ξ¯mjB̂ξ¯mjηÂξ¯mj+Âξ¯mjηB̂ξ¯mjB̂ξ¯mjηÂξ¯mj,E60

with η=etanhξ¯mj.

Then, the expectation value of Kx for the TMS number state of Eq. (55) can be given by

Kx=mjdv2cosh2ξ¯mjGmjv2(vt)2η+ηÂξ¯mjÂξ¯mj+B̂ξ¯mjB̂ξ¯mj+1E61

with Âξ¯mjÂξ¯mj+B̂ξ¯mjB̂ξ¯mj+1=3. From Eq. (25) we get the maximum value vt2=4, thus

Kx=6mjdvcosh2ξ¯mjGmjv2η+ηE62

where, η+η=2cosϕtanhξ¯mj, which implies that Kx=0, if ϕ=π2, this problem can be avoided by applying a local phase shift [53], we obtain

Kx=12cosϕmjdvGmjv2coshξ¯mjsinhξ¯mj=6cosϕmjdvGmjv2sinh2ξ¯mj.E63

If we take ϕ=2n+1π,n=0,1,2,3,, then Kx will be maximized. This implies that Kx>0, for ϕ=,n=0,1,2,3,. Therefore, for the state in Eq. (55), the L.H.S. of Eq. (58) is equal to zero and its R.H.S. is positive. Hence, we can confirm the inseparability of the state in Eq. (55) from the violation of the condition in Eq. (58), means that we can take it as an indicator of quantifying the amount of the entanglement of squeezed states which can be detected by increasing Kx. In other words, we have found that Kx increases as the spectral order j increases, which indicates that the amount of entanglement of the state in Eq. (55) increases.

Advertisement

5. Summary and comments

To summarize, we obtain the velocity phase matching, which can be used for determining the length of nonlinear crystal in the experiment setup to produce X a wave with velocity v. We have introduced the relation between the squeezing parameter and the spectral order of X waves which shows that the maximal squeezing increases as the spectral order increases, and there exists an optimal velocity (i.e., axicon angle) that maximizes the amount of squeezing generated for each value of the spectral order. Moreover, we have constructed the squeeze operator of the two modes X wave, then we act by it on the two-particle state generated by SPDC process to obtain the squeezed form of it. Finally, we have used the criterion of the quadrature moment of the fourth order to verify the entanglement of the two modes squeezed state of X wave. We have observed that the entanglement of the state increases as the spectral order increases.

References

  1. 1. Ing Konstantin Meyl, Scalar Waves, Theory and Experiments. Available from: http://www.k-meyl.de/go/Primaerliteratur/Scalar-Waves.pdf
  2. 2. Zohuri B. Scalar Wave Driven Energy Applications. Cham, Switzerland: Springer Nature Switzerland AG; 2019
  3. 3. Hernandez-Figueroa HE, Zamboni-Rached M, Recami E. Localized Waves. New York: Wiley; 2008
  4. 4. Busch P. Journal of Physics A: Mathematical and General. 1999;32:6535-6546
  5. 5. Jaeger G. Journal of Physics: Conference Series. 2020;1638:012010
  6. 6. Adhikari P, Choi J. Acta Physica Polonica B. 2017;48:145-161
  7. 7. Noah Glennon, Anthony E. Mirasola, Nathan Musoke, Mark C. Neyrinck, Chanda Prescod-Weinstein 2023, arXiv:2301.13220v1
  8. 8. Schrödinger E. Discussion of probability relations between separated systems. Mathematical Proceedings of the Cambridge Philosophical Society. 1935;31:555
  9. 9. Bell JS. On the Einstein-Podolsky-Rosen paradox. Physics 1. 1964:195-200
  10. 10. Aspect A, Grangier P, Roger G. Experimental tests of realistic local theories via Bell’s theorem. Physical Review Letters. 1981;47:460-463
  11. 11. Bennett CH, Brassard G, Crépeau C, Jozsa R, Peres A, Wootters WK. Teleporting an unknown quantum state via dual classical and Einstein-Podolsky-Rosen channels. Physical Review Letters. 1993;70:1895
  12. 12. Ekert AK. Quantum cryptography based on Bell’s theorem. Physical Review Letters. 1991;67:661
  13. 13. Shor PW. Polynomial-time algorithms for prime factorization and discrete logarithms on a quantum computer. SIAM Journal on Computing. 1997;26:1484
  14. 14. Acín A, Latorre JI, Pascual P. Three-party entanglement from positronium. Physical Review A. 2001;63:042107
  15. 15. Banerjee S, Kumar Alok A, Srikanth R, Hiesmayr BC. A quantum-information theoretic analysis of three-flavor neutrino oscillations. European Physical Journal C: Particles and Fields. 2015;75:487
  16. 16. Bertlmann RA, Hiesmayr BC. Kaonic qubits. Quantum Information Processing. 2006;5:421
  17. 17. Hiesmayr BC. Limits of quantum information In weak interaction processes of hyperons. Scientific Reports. 2015;5:11591
  18. 18. Bernabeu J. T and CPT symmetries in entangled neutral meson systems. Journal of Physics: Conference Series. 2011;335:012011
  19. 19. Peschanski R, Seki S. Entanglement entropy of scattering particles. Physics Letters B. 2016;758:89
  20. 20. Kharzeev DE, Levin EM. Deep inelastic scattering as a probe of entanglement. Physical Review D. 2017;95:114008
  21. 21. Maldacena J. A model with cosmological Bell inequalities. Fortschritte der Physik. 2015;64:10-23
  22. 22. Nielsen MA, Chuang IL. Quantum Computation and Quantum Information. Cambridge: Cambridge University Press; 2000
  23. 23. Bennett CH, DiVincenzo DP. Nature (London). 2000;404:247
  24. 24. Cerf NJ, Leuchs G, Polzik ES, editors. Quantum Information with Continuous Variables of Atoms and Light. London: Imperial College Press; 2007
  25. 25. Braunstein SL, van Loock P. Reviews of Modern Physics. 2005;77:513
  26. 26. Adesso G, Illuminati F. Journal of Physics A: Mathematical and Theoretical. 2007;40:7821
  27. 27. Reid MD et al. Reviews of Modern Physics. 2009;81:1727
  28. 28. Drummond P, Hiliry M. The Quantum Theory of Nonlinear Optics. New York, USA: Cambridge University Press; 2014
  29. 29. Boyd RW, Jha A, Malik M, O’Sullivan C, Rodenburg B, Gauthier DJ. Proceedings of SPIE. 2011;7948:79480L
  30. 30. Mirhosseini M, Magaña-Loaiza OS, O’Sullivan MN, Rodenburg B, Malik M, Lavery MPJ, et al. New Journal of Physics. 2015;17:033033
  31. 31. Mirhosseini M, Rodenburg B, Malik M, Boyd R. Journal of Modern Optics. 2013;61:43
  32. 32. Hamadou Ibrahim A, Roux FS, McLaren M, Konrad T, Forbes A. Physical Review A. 2013;88:012312
  33. 33. Gröblacher S, Jennewein T, Vaziri A, Weihs G, Zeilinger A. New Journal of Physics. 2006;8:75
  34. 34. Marrucci L, Karimi E, Slussarenko S, Piccirillo B, Santamato E, Nagali E, et al. Journal of Optics. 2011;13:064001
  35. 35. Cardano F, Massa F, Qassim H, Karimi E, Slussarenko S, Paparo D, et al. Science Advances. 2015;1:1500087
  36. 36. Jackson JD. Classical Electrodynamics. 3rd ed. New York: Wiley; 1998
  37. 37. Gopaul C, Andrews R. New Journal of Physics. 2007;9:94
  38. 38. Tyler GA, Boyd RW. Optics Letters. 2009;34:142
  39. 39. Anguita JA, Neifeld MA, Vasic BV. Applied Optics. 2008;47:2414
  40. 40. Krenn M, Fickler R, Fink M, Handsteiner J, Malik M, Scheidl T, et al. New Journal of Physics. 2014;16:113028
  41. 41. Durnin J, Miceli JJ, Eberly JH. Physical Review Letters. 1987;58:1499
  42. 42. Lu J, Greenleaf JF. IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control. 1992;39:19
  43. 43. Conti C. Generation of entangled 3D localized quantum wave-packets via optical parametric amplification. 2004. arXiv:quant-ph/0309069
  44. 44. Ciattoni A, Conti C. Journal of the Optical Society of America B: Optical Physics. 2007;24:2195
  45. 45. Ornigotti M, Conti C, Szameit A. Physical Review Letters. 2015;115:100401
  46. 46. Ornigotti M, Di Mauro Villari L, Szameit A, Conti C. Physical Review A. 2017;95:011802(R)
  47. 47. Ornigotti M, Conti C, Szameit A. Journal of Optics. 2018;20:065201
  48. 48. Olver FWJ, Lozier DW, Boisvert RF, Clark CW. NIST Handbook of Mathematical Functions. Cambridge: Cambridge University Press; 2010
  49. 49. Messiah A. Quantum Mechanics. New York: Dover; 2014
  50. 50. Mandel L, Wolf E. Optical Coherence and Quantum Optics. Cambridge, UK: Cambridge University Press; 1995
  51. 51. Loudon R. The Quantum Theory of Light. 3rd ed. Oxford: Oxford University Press; 1997
  52. 52. Gerry C, Knight P. Introductory Quantum Optics. Cambridge, UK: Cambridge University Press; 2005
  53. 53. Namiki R. Physical Review A. 2012;85:062307

Written By

Ali Saif M. Hassan, Waleed S.A. Hasan and Mohamed A. Shukri

Submitted: 17 June 2023 Reviewed: 02 July 2023 Published: 24 August 2023