Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Weak-light vector rogue waves, breathers, and their Stern-Gerlach deflection via electromagnetically induced transparency

Open Access Open Access

Abstract

We propose a scheme for generating and manipulating vector (or two-component) optical rogue waves using Akhmediev and Kuznetsov-Ma breathers in a coherent atomic system with an M-type five-level configuration via electromagnetically induced transparency (EIT). We show that the propagation velocity of these nonlinear excitations can be reduced to 10−4 c and their generation power can be lowered to microwatts. We also show that the motion trajectories of the two polarization components in these excitations can be deflected significantly by using a transversal gradient magnetic field, similar to the Stern-Gerlach effect of an atomic beam. We find that the deflection angle can reach to 10−4 radian within the propagation distance of only several centimeters; at variance with the atomic Stern-Gerlach effect, the deflection angle can be made different for different polarization components and may be actively adjusted in a controllable way. The results obtained may have promising applications, including the precise measurement of gradient magnetic fields.

© 2017 Optical Society of America

1. Introduction

Rogue waves are unexpectedly large, spatially and temporally localized extreme wave events excited from background states [1–3], firstly discovered in deep oceans [4, 5]. Due to the interest in both fundamental research and practical applications, in recent years the study on rogue waves has received tremendous attention in many branches of science, including fluid dynamics [6–9], nonlinear optics [10–14], plasma physics [15], acoustics [16], Bose-Einstein condensates in ultracold quantum gases [17], and even financial physics [18]. More references about rogue waves can be found in the comprehensive reviews on recent developments in the area of localized structures in optical and matter-wave media [19,20].

The nonlinear Schrödinger (NLS) equation, which governs the evolution of the envelope of weakly nonlinear, dispersive wave packets, has been widely employed as a simple model for describing rogue wave phenomena [21]. However, due to the requirement for modeling a variety of complex systems, it becomes necessary to investigate the rogue wave phenomena beyond the framework of a single-component NLS equation. Recent developments have considered dissipative effects [22–24], higher-order nonlinear terms [25–27], and in particular the interaction between several field components [28–34]. The latter investigations have led to the discovery of multi-component rogue waves, described by coupled NLS equations.

In this work, we suggest a scheme to create vector (or two-component) optical rogue waves, Akhmediev and Kuznetsov-Ma breathers in a coherent atomic system of a M-type five-level configuration interacting resonantly with a probe laser field (with two polarization components) and two control laser fields. By means of electromagnetically induced transparency (EIT), we prove that the propagation velocity of these nonlinear optical excitations can be reduced to 10−4 c (c is the light speed in vacuum) and the generation power of them can be lowered to the level of microwatt. We also prove that, under the action of a transversal gradient magnetic field, the motion trajectories of the two polarization components of these nonlinear optical excitations may undergo a significant deflection, very similar to the Stern-Gerlach (SG) effect of an atomic beam carrying with a spin angular momentum. We find that, within the propagation distance only several centimeters, the deflection angle of these nonlinear excitations the can reach to 10−4 radian. At variance with the Stern-Gerlach effect of atomic beams, the deflection angle can be made different for different polarization components and may be actively manipulated in an easy and controllable way.

Before proceeding, we note that the study of the SG deflection for linear optical pulses via EIT was firstly reported by Karpa and Weitz [35]. Later on, some further extensions related to the SG effect of optical pulses in atomic systems were also carried out in [36–38]. However, different from these previous works, the topics we consider here are vector optical rogue waves and breathers, and their SG deflection and active manipulation. The results obtained here not only provide a method for obtaining novel, multi-component optical rogue waves and breathers with coherent atomic gases, but also are promising for practical applications, e.g., for the precise measurement of gradient magnetic fields.

The rest of the article is arranged as follows. In Sec. II, the theoretical model under study is described. In Sec. III, the nonlinear envelope equations governing the evolution of the two polarization components of electromagnetic field are derived by using a method of multiple-scales. In Sec. IV, weak-light vector Peregrine rogue waves, Akhmediev and Kuznetsov-Ma breathers are presented, and their group velocity and generation power are estimated. In Sec. V, the SG effect of the rogue waves and breathers is discussed. Finally, in the last section a summary of main results obtained in this work is presented.

2. Model

We start with considering an atomic gas with a M-type five-level configuration. A linearly polarized, pulsed probe laser field (with pulse duration τ0) Ep = Ep1 + Ep2 = (∊̂p1 + ∊̂+p2) exp[i(kp zωpt)] + c.c. drives respectively the transitions |3〉 ↔ |2〉 and |3〉 ↔ |4〉 by its σ (i.e. left-circular) polarization component Ep1 and σ+ (i.e. right-circular) polarization component Ep2; two π-polarized continuous-wave (CW) control laser fields Ec1 = c1 exp[i(kc1xωc1t)] + c.c. and Ec2 = c2 exp[i(kc2xωc2t)] + c.c. drive respectively the transitions |1〉 ↔ |2〉 and |5〉 ↔ |4〉 [see Fig. 1]. Here c.c. denotes complex conjugate; ^(x^iy^)/2 and p1 [^+(x^+iy^)/2 and p2] are respectively the unit vector and envelope of the σ (σ+) polarization component of the probe field; , ŷ and are respectively unit vectors along coordinate axes x, y, and z. The control field envelopes c1 and c2 are assumed to be strong enough so that they can be taken to be undepleted during the evolution of the probe field. Furthermore, the atoms are assumed to be initially prepared in the ground state |3〉 and trapped in a gas cell with ultracold temperature to cancel Doppler broadening and collisions. The level diagram in Fig. 1 can be taken to be composed by two Λ-type EIT configurations, sharing with the common ground state |3〉.

 figure: Fig. 1

Fig. 1 Possible geometrical arrangement and the coordinate frame chosen for the system. A cold atomic gas with a M-type five-level configuration (right side) are loaded in a gas cell (left side). A pulsed probe field Ep drives respectively the transitions |3〉 ↔ |2〉 and |3〉 ↔ |4〉 by its σ polarization component and σ+ polarization component; Two strong CW control field Ec1 and Ec2 drive respectively the transitions |1〉 ↔ |2〉 and |5〉 ↔ |4〉. δcj, δp, and δp + Δ are detunings. Atoms are assumed to be prepared in the ground-state |3〉, indicated by black dots. B is a magnetic field applied to the atomic gas.

Download Full Size | PDF

Under electric-dipole and rotating-wave approximations, the Hamiltonian of the system in interaction picture reads Hint/ħ = (δpδc1)|1〉 〈1| + δp|2〉 〈2| + (δp + Δ)|4〉 〈4| + (δp + Δ − δc2)|5〉 〈5|+Ωc1|2〉 〈1|+Ωp1|2〉 〈3|+Ωp2|4〉 〈3|+Ωc2|4〉 〈5|+H.c., where Ωp1=−(p23 · ∊̂)p1/ħ and Ωp2=−(p43 · ∊̂+)p2/ħc1=−(p21 · )c1/ħ and Ωc2=−(p45 · )c2/ħ] are Rabi frequencies of the two circularly polarized components of the probe field (the two π-polarized control fields), with pjl being the electric dipole matrix element associated with the transition from |j〉 to |l〉. Detunings are defined as δp = ω23 + μ23B0ωp, δc1 = ω21 + μ21B0ωc1, δc2 = ω45 + μ45B0ωc2, and Δ = μ42B0, where μjl=μB(gFjmFjgFlmFl)/ (μB, gFj, and mFj are Bohr magneton, gyromagnetic factor, and magnetic quantum number of the level |j〉, respectively) and ωjl = (EjEl)/ħ, with Ej the eigenenergy of the state |j〉. Here we have applied an homogeneous static magnetic field, B = B0, to the atomic gas, which results in a Zeeman level shift ΔEZeeman=μBgFjmFjB0, and hence removes the degeneracy of ground-state sublevels |j〉 (j = 1, 3, 5) and the excited-state sublevels |j〉 (j = 2, 4). The motion of the atoms is governed by the optical Bloch equation

(t+Γ)ρ=i[Hint,ρ],
where ρ is 5 × 5 density-matrix and Γ is a 5 × 5 relaxation matrix (representing spontaneous emission and dephasing) of the system. The explicit expression of Eq. (1) is presented in Appendix A.

The evolution of Ep is controlled by Maxwell equation ∇2Ep − (1/c2)Phys. Lett. 2Ep/Phys. Lett. t2 = (1/0c2)Phys. Lett. 2Pp/Phys. Lett. t2, where Pp is the electric polarization intensity of the probe field. Under slowly varying envelope approximation, the Maxwell equation reduces to

[i(z+1ct)+c2ωp2]Ωp1,p2κ32,34ρ23,43=0,
where 2=2/x2+2/y2 and κ32,34 = Na|p32,34 · ∊̂|2ωp/(2ħ∊0c), with 0 the vacuum dielectric constant and Na the atomic density. The terms c2ωp2Ωpj (j = 1, 2) in Eq. (2) describe the diffraction effect of Ωpj.

If the transverse size of the probe pulse is large enough so that the diffraction effect is negligible, the solution for the linear propagation of the probe field can be obtained by taking Ωpj as small quantities. Then, from the Maxwell-Bloch (MB) equations (1) and (2), one finds that Ωpj are proportional to exp{i[Kj (ω)zωt)]}, with the linear dispersion relation

K1,2(ω)=ωc+κ32,34ωd1,5D1,2,
with D1,2 = |Ωc1,c2|2 − (ωd1,5)(ωd2,4). Here d1 = (δpδc1) − iΓ1/2, d2 = δpiΓ2/2, d3 = −iΓ3/2, d4 = (δp + Δ) − iΓ4/2, and d5 = (δp + Δ − δc2) − iΓ5/2.

From Eq. (3) we see that the linear dispersion relation of the system has two independent branches (i.e. the branch K1 and the branch K2). Shown in Fig. 2(a) is the imaginary part Im(K1) and the real part Re(K1) for the branch K1, as functions of ω. The red dashed (solid) line in Fig. 2(a) is the profile of Im(K1) for Ωc1 = 0 (Ωc1 = 1.0 × 107 s−1); the black dashed-dotted line is the profile of Re(K1) for Ωc1 = 1.0 × 107 s−1. Shown in Fig. 2(b) is the same as Fig. 2(a), but for the branch K2. When plotting Fig. 2, system parameters are chosen from a laser-cooled 87Rb atomic gas with atomic states assigned as |1〉 = |52S1/2, F = 2, mF = −1〉, |2〉 = |52P1/2, F = 2, mF = −1〉, |3〉 = |52S1/2, F = 1, mF = 0〉, |4〉 = |52P1/2, F = 2, mF = 1〉, and |5〉 = |52S1/2, F = 2, mF = 1〉. The atomic state decay rates are given by Γ2 ≃ Γ4 ≃ 0.5×107 s−1 and Γ1 ≃ Γ3 ≃ Γ5 ≃ 1.0 × 104 s−1. In addition, we take κ32κ34 = 1.0 × 109 cms−1, Ωc1 = Ωc2 = 1.0 × 107 s−1, δp = δc1 = 0, δc2 = 2.0 × 106 s−1, and Δ = 2.0 × 106 s−1.

 figure: Fig. 2

Fig. 2 Linear dispersion relation of the system. (a) Im(K1) and Re(K1) as functions of ω. The red dashed (solid) line is Im(K1) for the case Ωc1 = 0 (Ωc1 = 1.0 × 107 s−1); the black dashed-dotted line is Re(K1) for Ωc1 = 1.0 × 107 s−1. (b) The same as (a) but for Im(K2) and Re(K2).

Download Full Size | PDF

From Fig. 2(a) and Fig. 2(b) we see that, in the presence of the control fields, a large and deep EIT transparency window is opened in both the absorption spectra of Ωp1 and Ωp2 (i.e. the two polarization components of the probe pulse). This is a kind of double EIT phenomenon and it is resulted from the quantum destruction interference contributed by the two control fields. Furthermore, due to the symmetry of the two Λ-type level configurations, the linear dispersion relations for Ωp1 and Ωp2 are nearly coincided with each other. Due to the EIT effect and the symmetry of level configuration, the group velocities of the two polarization components [defined by Vgj = Re(∂Kj/∂ω)−1, j = 1, 2] can be very small compared with c for large control fields and can be matched very well [see the expression (9) below]. We shall see that the ultraslow, matched group velocities are very crucial for obtaining significant optical Stern-Gerlach deflections of the vector rogue waves, as shown below.

3. Asymptotic expansion and nonlinear envelope equations

One of the main goals of the present work is to obtain vector rogue waves and breathers in the system described above. To this end, we employ the method of multiple-scales [37, 38] to derive nonlinear envelope equations of the envelopes of the two probe-field components. We take the following asymptotic expansion ρmm=l=1lρmm(l) (m = 1–5), ρ33=1+l=1lρ33(l), ρmn=l=1lρmn(l) (m, n = 1–5; mn), and Ωpj=l=1lΩpj(l) (j = 1, 2). Here is a small parameter characterizing the small population depletion of the ground state |3〉, and all quantities on the right hand side of the asymptotic expansion are considered to be functions of the multi-scale variables zl=l z (l = 0, 1, 2), x1= ∊x, and tl=lt (l = 0, 1).

Substituting the above expansion into the optical MB equations (1) and (2), we obtain a chain of linear, but inhomogeneous equations, which can be solved order by order. At the first order (l = 1), we obtain the solution for the half Rabi frequencies of the probe field Ωpj(1)=Fjexp{i[Kj(ω)z0ωt0]} (j = 1, 2), where Fj are yet to be determined envelope functions of slow variables z2 and t2, and Kj (ω) is given by the linear dispersion relation (3). The solution for the density-matrix elements at this order reads ρ13(1)=Ωc1*Ωp1(1)/D1, ρ23(1)=(ωd1)Ωp1(1)/D1, ρ34(1)=(ωd5*)Ωp2*(1)/D2*, and ρ35(1)=Ωc2Ωp2*(1)/D2*, with the other density-matrix elements being zero.

At the second order (l = 2), one obtains the solution for ρmm(1) (m = 1–5), ρ12(1), ρ14(1), ρ15(1) ρ24(1), ρ25(1), and ρ45(1), which are lengthy and hence omitted here. The other second-order density-matrix elements are zero. A solvability condition for Ωpj(2) requires

i(Fjz1+1VgjFjt1)=0,(j=1,2)
where
Vg1,g2={1c+κ32,34|Ωc1,c2|2+(ω+d1,5)2D1,22}1,
are group velocities of F1 and F2, respectively.

At the third order (l = 3), a solvability condition for Ωpj(3) yields

iF1z2K1222F1t12(W11|F1|2+W12|F2|2)e2α¯1z2F1=0,
iF2z2K2222F2t12(W21|F1|2+W22|F2|2)e2α¯2z2F2=0,
where
K12,22=2κ32,34d2,4|Ωc1,c2|2+2d1,5|Ωc1,c2|2+d1,53(|Ωc1,c2|2d2,4d1,5)3,
W11,22=κ32,34d1,5|d1,5|2+|Ωc1,c2|2D1,2|D1,2|2,
W12,21=κ32,34d1,5|d5,1|2+|Ωc2,c1|2D1,2|D2,1|2,
with j =−2Im[Kj (ω = 0)]. Here, K12 and K22 characterize, respectively, the group velocity dispersions of F1 and F2; W11,22 and W12,21 characterize, respectively, self-phase and cross-phase modulations of the two polarization components of the probe field.

For the convenience of following discussions, we introduce δ = (1/Vg1 − 1/Vg2)/2, Vg = 2Vg1Vg2/(Vg1 + Vg2), and τ = tz/Vg. Then Eq. (6) combined with Eq. (4) can be written into the dimensionless form

iu1s+igδu1σgD12u1σ2(g11|u1|2+g12|u2|2)u1=ia1u1,
iu2sigδu2σgD22u2σ2(g22|u2|2+g21|u1|2)u2=ia2u2,
with s = z/LD, σ=2τ/τ0, uj = (Ωpj/U0)e−Im[Kj|ω=0]z, gδ = sgn(δ)LD/Lδ, gD1 = K12/|K22|, and gD2 = sgn(K22). Here, LDτ02/|K22| and Lδ = τ0/|δ| are respectively the characteristic dispersion length and characteristic group velocity mismatch length, U0 is the typical Rabi frequency of the probe field, g11,12,21,22 = W11,12,21,22/|W22| are nonlinear coefficients characterizing respectively self-phase (g11,22) and cross-phase (g12,21) modulations, and aj=Im[Kj |ω=0]LD are small absorption coefficients contributed mainly by the decay rates Γ2 and Γ4. Note that we have assumed LD = LNL, where LNL is the characteristic nonlinearity length, defined by LNL1/(U02|W22|).

By taking realistic parameter set Ωc1 = Ωc2 = 1.6 × 108 s−1, δp = 1.0 × 108 s−1, δc1 = 0, δc2 = 3.0 × 106 s−1, and Δ = 2.0 × 106 s−1, with other parameters the same as those given in Sec. II, we obtain K12 = (−4.56 + i0.25) × 10−15 cm−1s2, K22 = (−4.64 + i0.26) × 10−15 cm−1s2, W11 ≈ (−9.37 + i0.15) × 10−16 cm−1s2, W12 ≈ (−9.44 + i0.15) × 10−16 cm−1s2, W21 ≈ (−9.34 + i0.15) × 10−16 cm−1s2, W22 ≈ (−9.40 + i0.15) × 10−16 cm−1s2. We see that the imaginary parts of these quantities are much smaller than their relevant real parts. The physical reason for such small imaginary parts is due to the EIT effect induced by two CW control fields. With these parameters, the numerical values of the group velocities of the two polarization components of the probe field (5) read

Vg12.28×104c,
Vg22.26×104c,
which are very slow compared with c, and are very well matched each other. When taking τ0 = 6.0 × 10−8s and U0 = 3.6 × 107 s−1, we get the dimensionless coefficients in Eq. (8), given by gδ ≈ 0.01, gD1gD2 ≈ −1, LD ≈ 0.8 cm, g11g12g21g22 ≈ −1, and a1a2 ≈ 0.08.

Since gδ and a1,2 are small, one can treat the terms igδ∂uj/∂σ and −iajuj as a small perturbation. Then Eq. (8) after neglecting the small perturbation reduces to

iu1s+2u1σ2+(|u1|2+|u2|2)u1=0,
iu2s+2u2σ2+(|u2|2+|u1|2)u2=0.
Notice that Eq. (10) has the form of Manakov equation, which is integrable and supports analytical soliton solutions [39].

4. Ultraslow weak-light vector rogue waves, breathers and their interactions

To present a simple solution of Eq. (10), we introduce the transform [34] (u1,u2)=(cosα+sinα,cosαsinα)Ψ/2. Here α is a free, real constant and Ψ obeys the standard (1+1)-dimensional NLS equation i∂Ψ/∂s + 2Ψ/∂σ2 + |Ψ|2Ψ = 0, which admits the Peregrine soliton solution with the form Ψ = [1−4(1+2is)/(1+4s2+2σ2)]eis. Then, we have the two-component Peregrine soliton (vector rogue wave) solution with the form

u1(s,σ)=12(cosα+sinα)(141+2is1+4s2+2σ2)eis,
u2(s,σ)=12(cosαsinα)(141+2is1+4s2+2σ2)eis,
which, after returning to original variables, reads
Ωp1(z,τ)=U02(cosα+sinα)[141+2i(z/LD)1+4(z/LD)2+2(2τ/τ0)2]eiz/LD,
Ωp2(z,τ)=U02(cosαsinα)[141+2i(z/LD)1+4(z/LD)2+2(2τ/τ0)2]eiz/LD,
Using the relation (Ωp1, Ωp1) ∝ (p1, p2) and substituting (12) into the expression of the probe field Ep = (∊̂p1 + ∊̂+p2) exp[i(kp zωpt)] + c.c., we see that the excitation in the probe field is a rogue wave with two (circular) polarization components.

The Manakov equation (10) also admits solutions of vector Kuznetsov-Ma breather and Ahkmediev breather. In our system, the (dimensional) vector Kuznetsov-Ma breather solution reads

ΩpKBM1(z,τ)(z,τ)=U02(14q)cos(az/LD)+2qcosh(Ωτ/τ0)iasin(az/LD)2qcosh(Ωτ/τ0)cos(az/LD)×(cosα+sinα)eiz/LD,
ΩpKBM2(z,τ)(z,τ)=U02(14q)cos(az/LD)+2qcosh(Ωτ/τ0)iasin(az/LD)2qcosh(Ωτ/τ0)cos(az/LD)×(cosαsinα)eiz/LD.
where q = (1+Ω2/4)/2 and a=8q(2q1), with Ω and q real parameters satisfying q > 1/2. The vector Ahkmediev breather is given by
ΩpAB1(z,τ)=U02(14q)cosh(az/LD)+2qcos(Ωτ/τ0)+iasinh(az/LD)2qcos(Ωτ/τ0)cosh(az/LD)
ΩpAB2(z,τ)=U02(14q)cosh(az/LD)+2qcos(Ωτ/τ0)+iasinh(az/LD)2qcos(Ωτ/τ0)cosh(az/LD)
where q = (1−Ω2/4)/2 and a=8q(12)q, with Ω and q real parameters satisfying q < 1/2. Notice that when q = 1/2, both the vector Kuznetsov-Ma breather and the Ahkmediev breather degenerate into the vector Peregrine soliton (12); more sophisticate vector rogue waves and breathers of the Manakov equation can be found by using the Darboux technique [19,29,40].

Shown in Fig. 3 are numerical results on the propagation of the vector optical rogue wave and breathers as functions of 2τ/τ0 and z/LD by using split-step Fourier method for α = π/3 and Ω=2. The left and right panels in the figure are the intensity distributions for the first (i.e. σ) and the second (i.e. σ+) polarization components of the probe field, respectively. The upper panels [(a) and (b) ] are the intensity distributions for the first and second polarization components of the vector Kuznetsov-Ma breather, i.e. |ΩpKMB1/U0|2 (left panel) and |ΩpKMB2/U0|2 (right panel), obtained by using the vector Kuznetsov-Ma breather (13) as an initial condition; the middle panels [(c) and (d) ] are the intensity distributions for the vector Peregrine soliton, i.e. |Ωp1/U0|2 (left panel) and |Ωp2/U0|2 (right panel), obtained by using the vector Kuznetsov-Ma breather (11) as an initial condition; the lower panels [(e) and (f) ] are the intensity distributions for the vector Ahkmediev breathers, i.e. |ΩpAB1/U0|2 (left panel) and |ΩpAB2/U0|2 (right panel), obtained by using the vector Kuznetsov-Ma breather (14) as an initial condition. From the figure we see that waveshapes of the two probe-field components are matched well, but the intensity of the first (left panel) polarization component is different from that of the second (right panel) polarization component.

 figure: Fig. 3

Fig. 3 Vector optical rogue waves and breathers as functions of 2τ/τ0 and z/LD. The left and right panels are the intensity distributions for the first (i.e. σ) and the second (i.e. σ+) polarization components of the probe field, respectively. (a) [(b)] |ΩpKMB1/U0|2 [|ΩpKMB2/U0|2] obtained by using the vector Kuznetsov-Ma breather (13) as an initial condition; (c) [(d)] |Ωp1/U0|2 (left) [|Ωp2/U0|2] obtained by using the vector Peregrine soliton (11) as an initial condition; (e) [(f)] |ΩpAB1/U0|2 [|ΩpAB2/U0|2] obtained by using the vector Ahkmediev breathers (14) as an initial condition.

Download Full Size | PDF

By using the relation Vg = 2Vg1Vg2/(Vg1 + Vg2) and the result (9), we obtain

Vg2.27×104c.
Because the vector optical rogue wave (12), Akhmediev and Kuznetsov-Ma breathers (13) and (14) depend on the combined variable τ = tz/Vg, from (15) we see that the optical rogue wave and breathers obtained above have a very slow propagation velocity.

The threshold of the optical power Pth for generating the vector rogue waves and breathers in the present system can be estimated by using the Poynting’s vector. Assuming that the beam diameter ≈ 0.1 mm, we have

Pth0.97μW.
Thus, very low input power is needed to generate such vector optical rogue waves and breathers. The reason is that the self-phase and cross-phase modulations (Kerr effect) of the system are greatly enhanced by the EIT effect induced by the two control fields. Based on the results by (15) and (16), we called the nonlinear optical excitations given by (12), (13) and (14) the ultraslow, weak-light vector rogue waves and breathers.

To explore the interaction property of the optical rogue waves and breathers obtained above, a numerical simulation on two-soliton collisions is carried out by taking the initial condition u′1,2(s = 0) = u1,2(σ + d)ei0.5σ + ρu1,2(σd)ei0.5σ+ with ρ = 1 and d = 5. The parameter φ is the initial relative phase between the two solitons. Fig. 4(a) [Fig. 4(b)] shows motion trajectories of two Peregrine solitons during their collision with φ = 0 (φ = π/4). We see that for φ = 0 (φ = π/4) the interaction between the two solitons is attractive (repulsive). The physical reason is that for φ = 0 (φ = π/4) there is an increase (a decrease) of the optical refractive index when the two solitons approach each other, resulting in an attraction (a repulsion) of more photons into (from) the central region of the collision. In the figure, only the first polarization component is plotted; similar result for the second polarization component is also obtained but omitted here. Shown in Fig. 4(c) and Fig. 4(d) are respectively collisions between two Kuznetsov-Ma breathers for attractive (φ = 0) and repulsive (φ = π/4) interactions.

 figure: Fig. 4

Fig. 4 Collisions between two vector optical rogue waves and breathers. (a) ((b)) Motion trajectories of two Peregrine solitons with attractive (repulsive) interaction during collision; (c) ((d)) Motion trajectories of two Kuznetsov-Ma breathers with attractive (repulsive) interaction during collision. In panels (a) and (c), the initial phase between the two solitons φ = 0; in panels (b) and (d), φ = π/4.

Download Full Size | PDF

5. Stern-Gerlach effect of the Kuznetsov-Ma breathers

One of the main advantages of the present atomic system is that it is possible to realize an active control over the vector optical rogue waves and breathers by manipulating system parameters. As an example, we consider the inhomogeneous static magnetic field

B=z^B(x)=z^(B0+B1x),
is applied into the system (see Fig. 1). Here B1B0, B0 contributes to a Zeeman level shift and B1 is a transverse gradient magnetic field, which may result in a SG deflection of the polarization components of the probe field, as shown below. We assume the inhomogeneous static magnetic field can be expressed as B(x1)=B0 + 2 1x1, where 1 = −2B1O(B0). Then, the detunings of the system can be expanded as δp=δp(0)+2δp(2), Δ = Δ(0) + 2Δ(2), δc1=δc1(0)+2δc1(2), and δc2=δc2(0)+2δc2(2), where δp(0)=ω23+μ23B0ωp, Δ(0) = μ42B0, δc1(0)=ω21+μ21B0ωc1, δc2(0)=ω45+μ45B0ωc2, δp(2)=μ23B1x1, Δ(2) = μ42 B1 x1, δc1(2)=μ21B1x1, and δc2(2)=μ45B1x1. Note that the detunings at the leading order are homogeneous in space whereas they becomes inhomogeneous in space at the second order and are dependent on the gradient magnetic field B1.

The nonlinear envelope equation in the presence of the gradient magnetic field can also be derived by the use of the multiple-scale perturbation method, as done in the last section. But here we are interested in the case in which the probe-field envelope depends on the slow variables x1, z2, and t2 (i.e. the dispersion effect of the system is negligible, valid for the probe pulse with a large time duration). Carrying out the perturbation calculation to the third order, we obtain the following envelope equations:

i(z2+1Vg1t2)F1+c2ωp2F1x12(W11|F1|2+W12|F2|2)e2a¯1z2F1+M1B1x1F1=0,
i(z2+1Vg2t2)F2+c2ωp2F2x12(W22|F2|2+W21|F1|2)e2a¯2z2F2+M2B2x1F2=0,
with M1,2=κ32,34[d1,52μ23,43+|Ωc1,c2|2μ13,53]/D1,22, and other coefficients in the above equations the same as those given in the last section.

Equations (18) can be written into the dimensionless form

i(s+1λ1τ)u1+2u1ξ2(g11|u1|2+g12|u2|2)u1+1ξu1=ib1u1,
i(s+1λ2τ)u2+2u2ξ2(g22|u1|2+g21|u2|2)u2+2ξu2=ib2u2,
with 1,2=LDiffM1,2RB1/2 characterizing the gradient magnetic field. Here s = z/LDiff, τ = t/τ0, ξ=2x/R, λj = Vgjτ0/LDiff, and uj = (Ωpj/U0)e−Im[Kj|ω=0]z; LDiffωpR2/c and R are, respectively, the characteristic diffraction length and the radius of the probe beam; bj = Im[Kj |ω=0]LDiff are small absorption coefficients depending mainly on the decay rates Γ2 and Γ4. Note that we have assumed that the diffraction length is equal to the nonlinear length of the system, i.e., LDiff = LNL.

Due to the presence of the gradient magnetic field, the two polarization components of the probe pulse, u1 and u2, will separate each other after propagating some distance, and hence the cross-phase-modulation terms can be neglected. In this case, Eq. (19) reduces to

i(s+1λ1τ)u1+2u1ξ2g11|u1|2u1+1ξu1=ib1u1,
i(s+1λ2τ)u2+2u2ξ2g22|u1|2u1+2ξu2=ib2u2.
To solve the above equation, we assume u(s, τ, ξ)1,2 = G1,2(s, τ)v1,2(τ, ξ) with
G1,2(s,τ)=21/4g11,22e(sλ1,2τ)2/4=21/4g11,22e(zVg1,2t)2/(4LDiff2).
Then Eq. (20), after neglecting the small absorption coefficients bj, is further reduced into
iλ1v1τ+2v1ξ2|v1|2v1+1ξv1=0,
iλ2v2τ+2v2ξ2|v2|2v2+2ξv2=0.
Using the transformation v1,2=v1,2exp[i(1,2ξ1,2+1,22τ1,22/3)τ1,2] with ξ1,2=ξ1,2τ1,22/2 and τ′1,2 = λ1,2τ, Eq. (22) is simplified to iv1/τ1+2v1/ξ12+|v1|2v1=0 and iv2/τ2+2v2/ξ22+|v2|2v2=0, which support solutions of rogue waves and breathers, as shown in the last section.

In this way, we can get various solutions of vector rogue waves, Kuznetsov-Ma and Ahkmediev breathers of Eq. (20) in the presence of the SG gradient magnetic field. In the following, we only discuss the solution of vector Kuznetsov-Ma breather, which reads

u1(s,τ,ξ)=(14q)cos(aλ1τ)+2qcosh[2Ω(ξ1λ12τ2/2)]iasin(aλ1τ)2qcosh[2Ω(ξ1λ12τ2/2)]cos(aλ1τ)×21/4g11(cosα+sinα)eiλ1τ+i1λ(ξ1λ12τ2/6)τ(sλ1τ)2/4,
u2(s,τ,ξ)=(14q)cos(aλ2τ)+2qcosh[2Ω(ξ2λ22τ2/2)]iasin(aλ2τ)2qcosh[2Ω(ξ2λ22τ2/2)]cos(aλ2τ)×21/4g22(cosα+sinα)eiλ2τ+i2λ(ξ2λ22τ2/6)τ(sλ2τ)2/4,
where q = (1+Ω2/4)/2 and a=8q(2q1) (Ω and q are real parameters satisfying q > 1/2). Returning to original variables, we obtain the half Rabi frequencies of the two polarization components of the probe pulse
ΩpKBM1(z,t)=U0(14q)cos(aλ1t/τ0)+2qcosh[2Ω(x/R1λ12t2/(2τ02))]iasin(aλ1t/τ0)2qcosh[2Ω(x/R1λ12t2/(2τ02))]cos(aλ1t/τ0)×21/4g11(cosα+sinα)eiλ1t/τ0+i1λ1[x/R1λ12t2/(6τ02)]t/τ0(zVg1t)2/(4LDiff2),
ΩpKBM2(z,t)=U0(14q)cos(aλ2t/τ0)+2qcosh[2Ω(x/R2λ22t2/(2τ02))]iasin(aλ2t/τ0)2qcosh[2Ω(x/R2λ22t2/(2τ02))]cos(aλ2t/τ0)×21/4g22(cosαsinα)eiλ2t/τ0+i2λ2[x/R2λ22t2/(6τ02)]t/τ0(zVg2t)2/(4LDiff2).
From the solution (24), we see that in the presence of the magnetic field gradient, i.e. B1 ≠ 0 (and hence 1,2 ≠ 0), the motion of the center position of the jth polarization component (j = 1, 2) of the vector Kuznetsov-Ma breather follows the following parabolic trajectory
(x,y,z)=(Rjλj22τ02t2,0,Vgjt).
Since the magnetic-field parameter j can be positive and negative, the parabolic trajectories for the first and second polarization components may be different, and hence a SG-like effect can occur for the motion of the breather.

To illustrate clearly the feature of the trajectory deflection of the vector Kuznetsov-Ma breather, a numerical simulation is carried out on Eq. (20) with a nonzero B1. System parameters are chosen as Ωc1 = Ωc2 = 1.6 × 108 s−1, δp = 1.0 × 108 s−1, δc1 = 0, δc2 = 3.0 × 106 s−1, Δ = 1.8 × 106 s−1, R = 57 μm, U0 = 1.72 × 107 s−1, and B1 = 2.2 × 10−3 G cm−1. We get LDiff ≈ 2.6 cm, 1 = 0.01, and 2 = −0.01. Shown in Fig. 5(a) and Fig. 5(c) are respectively the deflection trajectories of the first and second polarization components (i.e. |Ωp1/U0|2 and |Ωp2/U0|2) as functions of x/R and z/LDiff. Their corresponding contour maps (i.e. the motion trajectories in the xz plane) are plotted in Fig. 5(b) and Fig. 5(d). In the simulation, the initial condition is chosen as the Kuznetsov-Ma breather (24), with α = π/3 and Ω=2. From the figure, we see that the motion trajectories of the two polarization components of the breather are parabolic and symmetric with respect to the x axis, which can be regarded as a SG effect of optical rogue wave, very similar to the SG effect of an atomic beam. In fact, a close analog exists between the well-known SG effect for atomic beam described in textbook [41] and the SG effect for the vector Kuznetsov-Ma breather described here. An atom in the lowest S-state has a spin angular momentum and can display a SG effect when passing through a gradient magnetic field. The vector Kuznetsov-Ma breather considering here has also a spin-like angular momentum (i.e. the two polarization components) and hence may undergo a SG-like effect in the presence of the gradient magnetic field.

 figure: Fig. 5

Fig. 5 Stern-Gerlach deflection of the vector Kuznetsov-Ma breather in the presence of a gradient magnetic field. (a) Intensity distribution of the first polarization component of the probe field, i.e. |ΩpKMB1/U0|2, as a function of t/τ0 and z/LDiff, with (b) the corresponding contour map (motion trajectory in the xz plane); (c) Intensity distribution of the second polarization component of the probe field, i.e. |ΩpKMB2/U0|2, with (d) the corresponding contour map (motion trajectory in the xz plane). The initial condition is chosen to be the Kuznetsov-Ma breather (24), with α = π/3 and Ω=2.

Download Full Size | PDF

However, different from the atomic SG effect where the motion trajectories of atoms are always symmetric for two different spin components [41], in the SG deflection of the vector Kuznetsov-Ma breather the motion trajectories of the two polarization components can be asymmetric. The reason is that the two polarization components of the breather can propagate with different velocities. To show this we repeat the simulation by taking Ωc2 = 1.54 × 108 without changing other system parameters. Then the motion velocities of the two polarization components become (Vg1, Vg2) = (2.3, 1.9) × 10−4 c. In this situation the first polarization component of the breather keeps the same waveshape and trajectory as that shown in Fig. 5(a) and Fig. 5(b) [see Fig. 6(a) and 6(b)], whereas the second component displays a trajectory different from that shown in Fig. 5(c) and Fig. 5(d) [see Fig. 6(c) and Fig. 6(d)].

 figure: Fig. 6

Fig. 6 Stern-Gerlach deflection of the Kuznetsov-Ma breather in the presence of a gradient magnetic field. (a) Intensity distribution of the first polarization component of probe field, i.e. |ΩpKMB1/U0|2 versus 2x/R and z/LDiff, with (b) the corresponding contour map; (c)Intensity distribution of the second polarization component of probe field, i.e. |ΩpKMB2/U0|2, with (d) the corresponding contour map. Due to the different velocities, the deflections of the two polarization components are asymmetric.

Download Full Size | PDF

Equation (25) tells us that the larger the gradient magnetic field B1 (and hence the larger the value of 1,2) is, the more apparently the motion trajectory of the breather bends. To confirm this prediction, in Fig. 7(a) we show the numerical result on the symmetrical and asymmetrical SG deflections of the Kuznetsov-Ma breather, where the SG deflection distance x/R is taken as a function of z/LDiff. The blue lines are for B1 = 2.2 × 10−3 G cm−1 and the red lines are for B1 = 4.4 × 10−3 G cm−1. The upbent lines are for the first (i.e. σ) polarization component, and downbent lines are for the second (i.e. σ+) polarization component. The solid lines (dashed lines) are for the symmetrical (asymmetrical) SG deflections, realized by taking Ωc2 to be 1.6 × 108 s−1 (1.54 × 108 s−1). From the figure, we can see clearly that the SG deflection distance of the Kuznetsov-Ma breather becomes larger when the magnetic field gradient B1 increases.

 figure: Fig. 7

Fig. 7 Symmetrical and asymmetrical SG deflections of the Kuznetsov-Ma breather. (a) SG deflection distance x/R as a function of z/LDiff for B1 = 2.2 × 10−3 G cm−1 (blue lines) and 4.4 × 10−3 G cm−1 (red lines). The upbent lines are for the first (i.e. σ) polarization component, and downbent lines are for the second (i.e. σ+) polarization component. The solid lines (dashed lines) are for the symmetrical (asymmetrical) SG deflections, obtained by taking Ωc2 to be 1.6 × 108 s−1 (1.54 × 108 s−1). (b) SG deflection angles θ as a function of the magnetic field gradient B1 when the breather propagating to z = 4LDiff ≈ 10.4 cm. The blue (red) line is for σ (σ+) polarization component.

Download Full Size | PDF

One can make an estimation on the SG deflection angle. The expected deflection angle of the output jth polarization component of the Kuznetsov-Ma breather is determined by the ratio between the propagating velocity along the x axis (i.e. Vj=Rjλj2t/τ02) and that along the z axis (i.e. Vgj). Thus, after passing through the medium with length L the deflection angle of the output jth polarization component of the breather is given by

θj=VjVgj=MjR22LLDiffB1
(j = 1, 2). Fig. (7)(c) shows the deflection angles of the σ polarization component (blue line) and the σ+ polarization component (red line) as functions of the magnetic field B1. We see that the deflection angle can reach to 10−4 radian when the magnetic field B1 = 2.6 × 10−3 G cm−1 (B1 = 2.1 × 10−3 G cm−1) for σ (σ+) polarization component propagating to z = 4LDiff ≈ 10.4 cm.

The SG effect of the vector Kuznetsov-Ma breathers is one of examples illustrating how to actively manipulating the vector optical rogue waves and breathers. The reason for such active manipulation possible is due to the resonant characters of the multi-level atomic systems coupling with laser fields, in which many system parameters can be selected and adjusted in a controllable way. Besides the interest in fundamental research, such active manipulation maybe find promising practical applications. For instance, one can utilize the SG effect of the vector Kuznetsov-Ma breathers (or other rouge wave and breathers) to precisely detect the gradient magnetic field by measuring their SG deflection angles.

6. Conclusion

In this work, we have presented a scheme for creating vector optical rogue waves and breathers in a M-type five-level atomic system coupled with a probe laser field of two polarization components and two control laser fields. By means of EIT we have proved that the propagation velocity of these nonlinear optical excitations may be reduced to 10−4 c and their generation power may be lowered to the level of microwatt. We have also proved that, by the use of a transversal gradient magnetic field, the motion trajectories of the two polarization components of these nonlinear excitations can display a large deflection, a phenomenon quite similar to the SG effect of an atomic beam carrying with a spin angular momentum. We have found that within the propagation distance only several centimeters the deflection angle of these nonlinear excitations can reach to the value of 10−4 radian. At variance with the SG effect of atomic beams, the deflection angle can be made different for different polarization components and may be actively manipulated in an easy and controllable way. The results obtained here have not only provided a method for generating novel, multi-component optical rogue waves and breathers with coherent atomic gases at weak light level, but also opened an avenue for their active manipulation and possible practical applications in future.

A. The explicit expression of equation (1)

The explicit form of the optical Bloch equation (1), i.e., the equation of motion of the density-matrix elements ρjl (j, l = 1–5), reads

ρ11t=iΩc1*ρ21+iΩc1ρ12+Γ41ρ44+Γ21ρ22,
ρ22t=iΩc1ρ12+iΩc1*ρ21+iΩp1*ρ23iΩp1ρ32Γ2ρ22,
ρ33t=iΩp1*ρ23+iΩp1ρ32+iΩp2ρ34iΩp2*ρ43+Γ43ρ44+Γ23ρ22,
ρ44t=iΩp2ρ34+iΩp2*ρ43+iΩc2*ρ45iΩc2ρ54Γ4ρ44,
ρ55t=iΩc2*ρ45+iΩc2ρ54+Γ45ρ44+Γ25ρ22,
for the diagonal elements, and
ρ12t=iδc1ρ12iΩc1*(ρ22ρ11)+iΩp1*σ13Γ2+γ122ρ12,
ρ13t=i(δc1δp)ρ13+iΩp1ρ12+iΩp2ρ14iΩc1*ρ23γ132ρ13,
ρ14t=i(Δ+δc1)ρ14+iΩp2*ρ13+iΩc2*ρ15iΩc1*ρ24Γ4+γ142ρ14,
ρ15t=i(Δ+δc1δc2)ρ15+iΩc2ρ14iΩc1*ρ25γ152ρ15,
ρ23t=iδpρ23iΩc1ρ13+iΩp2ρ24iΩp1(ρ33ρ22)Γ2+γ232ρ23,
ρ24t=iΔρ24iΩc1ρ14+iΩc2*ρ25+iΩp2*ρ23iΩp1ρ34Γ2+Γ4+γ242ρ24,
ρ25t=i(Δδc2)ρ25iΩc1ρ15iΩp1ρ35+iΩc2ρ24Γ2+γ252ρ25,
ρ34t=i(δp+Δ)ρ34iΩp1*ρ24+iΩc2*ρ35iΩp2*(ρ44ρ33)Γ4+γ342ρ34,
ρ35t=i(δp+Δδc2)ρ35iΩp1*ρ25iΩp2*ρ45+iΩc2ρ34γ352ρ35,
ρ45t=iδc2ρ45iΩc2(ρ55ρ44)iΩp2ρ35Γ4+γ452ρ45,
for the nondiagonal elements. Here Γ2 = Γ12 + Γ32 + Γ52 and Γ4 = Γ14 + Γ34 + Γ54 are atomic decay rates, with Γjl representing the spontaneous emission decay rate from state |l〉 to state |j〉; γjl=(Γj+Γl)/2+γjldep, with γjldep the dephasing rates related to states |j〉 and |l〉 [42].

Funding

National Natural Science Foundation of China (11475063, 11474099); National Basic Research Program of China (2016YFA0302103).

References and links

1. M. Onorato, S. Residori, U. Bortolozzo, A. Montinad, and F. T. Arecchi, “Rogue waves and their generating mechanisms in different physical contexts,” Phys. Reports 528, 47 (2013). [CrossRef]  

2. N. Akhmediev, J. M. Dudley, D. R. Solli, and S. K. Turitsyn, “Recent progress in investigating optical rogue waves,” J. Opt. 15, 060201 (2013). [CrossRef]  

3. J. M. Dudley, F. Dias, M. Erkintalo, and G. Genty, “Instabilities, breathers and rogue waves in optics,” Nat. Photon. 8, 755 (2014). [CrossRef]  

4. C. Kharif, E. Pelinovsky, and A. Slunyaev, Rogue waves in the ocean (Springer, Heidelberg, 2009).

5. A. R. Osborne, Nonlinear Ocean Waves (Academic Press, 2009).

6. M. Shatz, H. Punzmann, and H. Xia, “Capillary rogue waves,” Phys. Rev. Lett 104, 104503 (2010). [CrossRef]  

7. A. Chabchoub, N. P. Hoffmann, and N. Akhmediev, “Rogue Wave Observation in a Water Wave Tank,” Phys. Rev. Lett. 106, 204502 (2011). [CrossRef]   [PubMed]  

8. A. Chabchoub, N. Hoffmann, M. Onorato, and N. Akhmediev, “Super Rogue Waves: Observation of a Higher-Order Breather in Water Waves,” Phys. Rev. X 2, 011015 (2012).

9. A. Chabchoub, N. Hoffmann, M. Onorato, A. Slunyaev, A. Sergeeva, E. Pelinovsky, and N. Akhmediev, “Observation of a hierarchy of up to fifth-order rogue waves in a water tank,” Phys. Rev. E 86, 056601 (2012). [CrossRef]  

10. D. R. Solli, C. Ropers, P. Koonath, and B. Jalali, “Optical rogue waves,” Nature (London) 450, 1054 (2007). [CrossRef]  

11. A. Montina, U. Bortolozzo, S. Residori, and F. T. Arecchi, “Non-Gaussian Statistics and Extreme Waves in a Nonlinear Optical Cavity,” Phys. Rev. Lett. 103, 173901 (2009). [CrossRef]   [PubMed]  

12. B. Kibler, J. Fatome, C. Finot, G. Millot, F. Dias, G. Genty, N. Akhmediev, and J. M. Dudley, “The Peregrine soliton in nonlinear fibre optics,” Nat. Phys. 6, 790 (2010). [CrossRef]  

13. A. Zaviyalov, O. Egorov, R. Iliew, and F. Lederer, “Rogue waves in mode-locked fiber lasers,” Phys. Rev. A 85, 013828 (2012). [CrossRef]  

14. J. Liu, C. Hang, and G. Huang, “Weak-light rogue waves, breathers, and their active control in a cold atomic gas via electromagnetically induced transparency,” Phys. Rev. A 93, 063836 (2016). [CrossRef]  

15. H. Bailung, S. K. Sharma, and Y. Nakamura, “Observation of Peregrine solitons in a multicomponent plasma with negative ions,” Phys. Rev. Lett. 107, 255005 (2011). [CrossRef]  

16. A. N. Ganshin, V. B. Efimov, G. V. Kolmakov, L. P. Mezhov-Deglin, and P. V. E. McClintock, “Observation of an inverse energy cascade in developed acoustic turbulence in superfluid helium,” Phys. Rev. Lett. 101, 065303 (2008). [CrossRef]   [PubMed]  

17. Yu. V. Bludov, V. V. Konotop, and N. Akhmediev, “Matter rogue waves,” Phys. Rev. A 80, 033610 (2009). [CrossRef]  

18. Z. Yan, “Vector financial rogue waves,” Phys. Lett. A 375, 4274 (2011). [CrossRef]  

19. D. Mihalache, “Localized structures in nonlinear optical media: A selection of recent studies”, Rom. Rep. Phys. 67, 1383 (2015).

20. D. Mihalache, “Multidimensional localized structures in optical and matter-wave media: A topical survey of recent literature”, Rom. Rep. Phys. 69, 403 (2017).

21. N. Akhmediev, J. M. Soto-Crespo, and A. Ankiewicz, “Extreme waves that appear from nowhere: On the nature of rogue waves,” Phys. Lett. A 373, 2137–2145 (2009). [CrossRef]  

22. C. Lecaplain, Ph. Grelu, J. M. Soto-Crespo, and N. Akhmediev, “Dissipative Rogue waves generated by chaotic pulse bunching in a mode-locked laser,” Phys. Rev. Lett. 108, 233901 (2012). [CrossRef]   [PubMed]  

23. J. M. Soto-Crespo, Ph. Grelu, and N. Akhmediev, “Dissipative rogue waves: Extreme pulses generated by passively mode-locked lasers,” Phys. Rev. E 84, 016604 (2011). [CrossRef]  

24. Ph. Grelu and N. Akhmediev, “Dissipative solitons for mode-locked lasers,” Nat. Photon. 6, 84 (2012). [CrossRef]  

25. A. Ankiewicz, J. M. Soto-Crespo, and N. Akhmediev, “Rogue waves and rational solutions of the Hirota equation,” Phys. Rev. E 81, 046602 (2010). [CrossRef]  

26. U. Bandelow and N. Akhmediev, “Sasa-Satsuma equation: Soliton on a background and its limiting cases,” Phys. Rev. E 86, 026606 (2012). [CrossRef]  

27. S. Chen, “Twisted rogue-wave pairs in the Sasa-Satsuma equation,” Phys. Rev. E 88, 023202 (2013). [CrossRef]  

28. Y. Ohta and J. Yang, “Rogue waves in the Davey-Stewartson I equation,” Phys. Rev. E 86, 036604 (2012). [CrossRef]  

29. F. Baronio, A. Degasperis, M. Conforti, and S. Wabnitz, “Solutions of the vector nonlinear Schrödinger equations: Evidence for deterministic Rogue waves,” Phys. Rev. Lett. 109, 044102 (2012). [CrossRef]  

30. S. Chen and L.-Y. Song, “Rogue waves in coupled Hirota systems,” Phys. Rev. E 87, 032910 (2013). [CrossRef]  

31. K. W. Chow, H. N. Chan, D. J. Kedziora, and R. H. J. Grimshaw, “Rogue wave modes for the long wave-short wave resonance model,” J. Phys. Soc. Jpn. 82, 074001 (2013). [CrossRef]  

32. F. Baronio, M. Conforti, A. Degasperis, and S. Lombardo, “Rogue waves emerging from the resonant interaction of three waves,” Phys. Rev. Lett. 111, 114101 (2013). [CrossRef]   [PubMed]  

33. S. Chen, Ph. Grelu, and J. M. Soto-Crespo, “Dark- and bright-rogue-wave solutions for media with long-wave–short-wave resonance,” Phys. Rev. E 89, 011201(R) (2014). [CrossRef]  

34. Yu. V. Bludov, V. V. Konotop, and N. Akhmediev, “Vector rogue waves in binary mixtures of Bose-Einstein condensates,” Eur. Phys. J. Special Topics. 185, 169 (2010). [CrossRef]  

35. L. Karpa and M. Weitz, “A Stern-Gerlach experiment for slow light,” Nat. Phys. 2, 332 (2006). [CrossRef]  

36. Y. Guo, L. Zhou, L-M. Kuang, and C. P. Sun, “Magneto-optical Stern-Gerlach effect in an atomic ensemble,” Phys. Rev. A 78, 013833 (2008). [CrossRef]  

37. C. Hang and G. Huang, “Stern-Gerlach effect of weak-light ultraslow vector solitons,” Phys. Rev. A 86, 043809 (2012). [CrossRef]  

38. Z. Chen and G. Huang, “Stern-Gerlach effect of multi-component ultraslow optical solitons via electromagnetically induced transparency,” J. Opt. Soc. Am. 30, 2248 (2013). [CrossRef]  

39. S. V. Manakov, “On the theory of two-dimensional stationary selffocusing of electromagnetic waves,” Zh. Eksp. Teor. Fiz. 65, 505 (1973) [Sov. Phys. JETP 38, 248 (1974)].

40. S. Chen and D. Mihalache, “Vector rogue waves in the Manakov system: diversity and compossibility,” J. Phys. A: Math. Theor. 48, 215202 (2015). [CrossRef]  

41. C. J. Foot, Atomic Physics (Oxford University, 2005).

42. R. W. Boyd, Nonlinear Optics, 3rd ed. (Academic, Elsevier, 2008).

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 Possible geometrical arrangement and the coordinate frame chosen for the system. A cold atomic gas with a M-type five-level configuration (right side) are loaded in a gas cell (left side). A pulsed probe field E p drives respectively the transitions |3〉 ↔ |2〉 and |3〉 ↔ |4〉 by its σ polarization component and σ+ polarization component; Two strong CW control field Ec1 and Ec2 drive respectively the transitions |1〉 ↔ |2〉 and |5〉 ↔ |4〉. δcj, δp, and δp + Δ are detunings. Atoms are assumed to be prepared in the ground-state |3〉, indicated by black dots. B is a magnetic field applied to the atomic gas.
Fig. 2
Fig. 2 Linear dispersion relation of the system. (a) Im(K1) and Re(K1) as functions of ω. The red dashed (solid) line is Im(K1) for the case Ωc1 = 0 (Ωc1 = 1.0 × 107 s−1); the black dashed-dotted line is Re(K1) for Ωc1 = 1.0 × 107 s−1. (b) The same as (a) but for Im(K2) and Re(K2).
Fig. 3
Fig. 3 Vector optical rogue waves and breathers as functions of 2 τ / τ 0 and z/LD. The left and right panels are the intensity distributions for the first (i.e. σ) and the second (i.e. σ+) polarization components of the probe field, respectively. (a) [(b)] |ΩpKMB1/U0|2 [|ΩpKMB2/U0|2] obtained by using the vector Kuznetsov-Ma breather (13) as an initial condition; (c) [(d)] |Ωp1/U0|2 (left) [|Ωp2/U0|2] obtained by using the vector Peregrine soliton (11) as an initial condition; (e) [(f)] |ΩpAB1/U0|2 [|ΩpAB2/U0|2] obtained by using the vector Ahkmediev breathers (14) as an initial condition.
Fig. 4
Fig. 4 Collisions between two vector optical rogue waves and breathers. (a) ((b)) Motion trajectories of two Peregrine solitons with attractive (repulsive) interaction during collision; (c) ((d)) Motion trajectories of two Kuznetsov-Ma breathers with attractive (repulsive) interaction during collision. In panels (a) and (c), the initial phase between the two solitons φ = 0; in panels (b) and (d), φ = π/4.
Fig. 5
Fig. 5 Stern-Gerlach deflection of the vector Kuznetsov-Ma breather in the presence of a gradient magnetic field. (a) Intensity distribution of the first polarization component of the probe field, i.e. |ΩpKMB1/U0|2, as a function of t/τ0 and z/LDiff, with (b) the corresponding contour map (motion trajectory in the xz plane); (c) Intensity distribution of the second polarization component of the probe field, i.e. |ΩpKMB2/U0|2, with (d) the corresponding contour map (motion trajectory in the xz plane). The initial condition is chosen to be the Kuznetsov-Ma breather (24), with α = π/3 and Ω = 2 .
Fig. 6
Fig. 6 Stern-Gerlach deflection of the Kuznetsov-Ma breather in the presence of a gradient magnetic field. (a) Intensity distribution of the first polarization component of probe field, i.e. |ΩpKMB1/U0|2 versus 2 x / R and z/LDiff, with (b) the corresponding contour map; (c)Intensity distribution of the second polarization component of probe field, i.e. |ΩpKMB2/U0|2, with (d) the corresponding contour map. Due to the different velocities, the deflections of the two polarization components are asymmetric.
Fig. 7
Fig. 7 Symmetrical and asymmetrical SG deflections of the Kuznetsov-Ma breather. (a) SG deflection distance x/R as a function of z/LDiff for B1 = 2.2 × 10−3 G cm−1 (blue lines) and 4.4 × 10−3 G cm−1 (red lines). The upbent lines are for the first (i.e. σ) polarization component, and downbent lines are for the second (i.e. σ+) polarization component. The solid lines (dashed lines) are for the symmetrical (asymmetrical) SG deflections, obtained by taking Ωc2 to be 1.6 × 108 s−1 (1.54 × 108 s−1). (b) SG deflection angles θ as a function of the magnetic field gradient B1 when the breather propagating to z = 4LDiff ≈ 10.4 cm. The blue (red) line is for σ (σ+) polarization component.

Equations (57)

Equations on this page are rendered with MathJax. Learn more.

( t + Γ ) ρ = i [ H int , ρ ] ,
[ i ( z + 1 c t ) + c 2 ω p 2 ] Ω p 1 , p 2 κ 32 , 34 ρ 23 , 43 = 0 ,
K 1 , 2 ( ω ) = ω c + κ 32 , 34 ω d 1 , 5 D 1 , 2 ,
i ( F j z 1 + 1 V g j F j t 1 ) = 0 , ( j = 1 , 2 )
V g 1 , g 2 = { 1 c + κ 32 , 34 | Ω c 1 , c 2 | 2 + ( ω + d 1 , 5 ) 2 D 1 , 2 2 } 1 ,
i F 1 z 2 K 12 2 2 F 1 t 1 2 ( W 11 | F 1 | 2 + W 12 | F 2 | 2 ) e 2 α ¯ 1 z 2 F 1 = 0 ,
i F 2 z 2 K 22 2 2 F 2 t 1 2 ( W 21 | F 1 | 2 + W 22 | F 2 | 2 ) e 2 α ¯ 2 z 2 F 2 = 0 ,
K 12 , 22 = 2 κ 32 , 34 d 2 , 4 | Ω c 1 , c 2 | 2 + 2 d 1 , 5 | Ω c 1 , c 2 | 2 + d 1 , 5 3 ( | Ω c 1 , c 2 | 2 d 2 , 4 d 1 , 5 ) 3 ,
W 11 , 22 = κ 32 , 34 d 1 , 5 | d 1 , 5 | 2 + | Ω c 1 , c 2 | 2 D 1 , 2 | D 1 , 2 | 2 ,
W 12 , 21 = κ 32 , 34 d 1 , 5 | d 5 , 1 | 2 + | Ω c 2 , c 1 | 2 D 1 , 2 | D 2 , 1 | 2 ,
i u 1 s + i g δ u 1 σ g D 1 2 u 1 σ 2 ( g 11 | u 1 | 2 + g 12 | u 2 | 2 ) u 1 = i a 1 u 1 ,
i u 2 s i g δ u 2 σ g D 2 2 u 2 σ 2 ( g 22 | u 2 | 2 + g 21 | u 1 | 2 ) u 2 = i a 2 u 2 ,
V g 1 2.28 × 10 4 c ,
V g 2 2.26 × 10 4 c ,
i u 1 s + 2 u 1 σ 2 + ( | u 1 | 2 + | u 2 | 2 ) u 1 = 0 ,
i u 2 s + 2 u 2 σ 2 + ( | u 2 | 2 + | u 1 | 2 ) u 2 = 0 .
u 1 ( s , σ ) = 1 2 ( cos α + sin α ) ( 1 4 1 + 2 i s 1 + 4 s 2 + 2 σ 2 ) e i s ,
u 2 ( s , σ ) = 1 2 ( cos α sin α ) ( 1 4 1 + 2 i s 1 + 4 s 2 + 2 σ 2 ) e i s ,
Ω p 1 ( z , τ ) = U 0 2 ( cos α + sin α ) [ 1 4 1 + 2 i ( z / L D ) 1 + 4 ( z / L D ) 2 + 2 ( 2 τ / τ 0 ) 2 ] e i z / L D ,
Ω p 2 ( z , τ ) = U 0 2 ( cos α sin α ) [ 1 4 1 + 2 i ( z / L D ) 1 + 4 ( z / L D ) 2 + 2 ( 2 τ / τ 0 ) 2 ] e i z / L D ,
Ω p K B M 1 ( z , τ ) ( z , τ ) = U 0 2 ( 1 4 q ) cos ( a z / L D ) + 2 q cosh ( Ω τ / τ 0 ) i a sin ( a z / L D ) 2 q cosh ( Ω τ / τ 0 ) cos ( a z / L D ) × ( cos α + sin α ) e i z / L D ,
Ω p K B M 2 ( z , τ ) ( z , τ ) = U 0 2 ( 1 4 q ) cos ( a z / L D ) + 2 q cosh ( Ω τ / τ 0 ) i a sin ( a z / L D ) 2 q cosh ( Ω τ / τ 0 ) cos ( a z / L D ) × ( cos α sin α ) e i z / L D .
Ω p A B 1 ( z , τ ) = U 0 2 ( 1 4 q ) cosh ( a z / L D ) + 2 q cos ( Ω τ / τ 0 ) + i a sinh ( a z / L D ) 2 q cos ( Ω τ / τ 0 ) cosh ( a z / L D )
Ω p A B 2 ( z , τ ) = U 0 2 ( 1 4 q ) cosh ( a z / L D ) + 2 q cos ( Ω τ / τ 0 ) + i a sinh ( a z / L D ) 2 q cos ( Ω τ / τ 0 ) cosh ( a z / L D )
V g 2.27 × 10 4 c .
P th 0.97 μ W .
B = z ^ B ( x ) = z ^ ( B 0 + B 1 x ) ,
i ( z 2 + 1 V g 1 t 2 ) F 1 + c 2 ω p 2 F 1 x 1 2 ( W 11 | F 1 | 2 + W 12 | F 2 | 2 ) e 2 a ¯ 1 z 2 F 1 + M 1 B 1 x 1 F 1 = 0 ,
i ( z 2 + 1 V g 2 t 2 ) F 2 + c 2 ω p 2 F 2 x 1 2 ( W 22 | F 2 | 2 + W 21 | F 1 | 2 ) e 2 a ¯ 2 z 2 F 2 + M 2 B 2 x 1 F 2 = 0 ,
i ( s + 1 λ 1 τ ) u 1 + 2 u 1 ξ 2 ( g 11 | u 1 | 2 + g 12 | u 2 | 2 ) u 1 + 1 ξ u 1 = i b 1 u 1 ,
i ( s + 1 λ 2 τ ) u 2 + 2 u 2 ξ 2 ( g 22 | u 1 | 2 + g 21 | u 2 | 2 ) u 2 + 2 ξ u 2 = i b 2 u 2 ,
i ( s + 1 λ 1 τ ) u 1 + 2 u 1 ξ 2 g 11 | u 1 | 2 u 1 + 1 ξ u 1 = i b 1 u 1 ,
i ( s + 1 λ 2 τ ) u 2 + 2 u 2 ξ 2 g 22 | u 1 | 2 u 1 + 2 ξ u 2 = i b 2 u 2 .
G 1 , 2 ( s , τ ) = 2 1 / 4 g 11 , 22 e ( s λ 1 , 2 τ ) 2 / 4 = 2 1 / 4 g 11 , 22 e ( z V g 1 , 2 t ) 2 / ( 4 L Diff 2 ) .
i λ 1 v 1 τ + 2 v 1 ξ 2 | v 1 | 2 v 1 + 1 ξ v 1 = 0 ,
i λ 2 v 2 τ + 2 v 2 ξ 2 | v 2 | 2 v 2 + 2 ξ v 2 = 0 .
u 1 ( s , τ , ξ ) = ( 1 4 q ) cos ( a λ 1 τ ) + 2 q cosh [ 2 Ω ( ξ 1 λ 1 2 τ 2 / 2 ) ] i a sin ( a λ 1 τ ) 2 q cosh [ 2 Ω ( ξ 1 λ 1 2 τ 2 / 2 ) ] cos ( a λ 1 τ ) × 2 1 / 4 g 11 ( cos α + sin α ) e i λ 1 τ + i 1 λ ( ξ 1 λ 1 2 τ 2 / 6 ) τ ( s λ 1 τ ) 2 / 4 ,
u 2 ( s , τ , ξ ) = ( 1 4 q ) cos ( a λ 2 τ ) + 2 q cosh [ 2 Ω ( ξ 2 λ 2 2 τ 2 / 2 ) ] i a sin ( a λ 2 τ ) 2 q cosh [ 2 Ω ( ξ 2 λ 2 2 τ 2 / 2 ) ] cos ( a λ 2 τ ) × 2 1 / 4 g 22 ( cos α + sin α ) e i λ 2 τ + i 2 λ ( ξ 2 λ 2 2 τ 2 / 6 ) τ ( s λ 2 τ ) 2 / 4 ,
Ω p K B M 1 ( z , t ) = U 0 ( 1 4 q ) cos ( a λ 1 t / τ 0 ) + 2 q cosh [ 2 Ω ( x / R 1 λ 1 2 t 2 / ( 2 τ 0 2 ) ) ] i a sin ( a λ 1 t / τ 0 ) 2 q cosh [ 2 Ω ( x / R 1 λ 1 2 t 2 / ( 2 τ 0 2 ) ) ] cos ( a λ 1 t / τ 0 ) × 2 1 / 4 g 11 ( cos α + sin α ) e i λ 1 t / τ 0 + i 1 λ 1 [ x / R 1 λ 1 2 t 2 / ( 6 τ 0 2 ) ] t / τ 0 ( z V g 1 t ) 2 / ( 4 L Diff 2 ) ,
Ω p K B M 2 ( z , t ) = U 0 ( 1 4 q ) cos ( a λ 2 t / τ 0 ) + 2 q cosh [ 2 Ω ( x / R 2 λ 2 2 t 2 / ( 2 τ 0 2 ) ) ] i a sin ( a λ 2 t / τ 0 ) 2 q cosh [ 2 Ω ( x / R 2 λ 2 2 t 2 / ( 2 τ 0 2 ) ) ] cos ( a λ 2 t / τ 0 ) × 2 1 / 4 g 22 ( cos α sin α ) e i λ 2 t / τ 0 + i 2 λ 2 [ x / R 2 λ 2 2 t 2 / ( 6 τ 0 2 ) ] t / τ 0 ( z V g 2 t ) 2 / ( 4 L Diff 2 ) .
( x , y , z ) = ( R j λ j 2 2 τ 0 2 t 2 , 0 , V g j t ) .
θ j = V j V g j = M j R 2 2 L L Diff B 1
ρ 11 t = i Ω c 1 * ρ 21 + i Ω c 1 ρ 12 + Γ 41 ρ 44 + Γ 21 ρ 22 ,
ρ 22 t = i Ω c 1 ρ 12 + i Ω c 1 * ρ 21 + i Ω p 1 * ρ 23 i Ω p 1 ρ 32 Γ 2 ρ 22 ,
ρ 33 t = i Ω p 1 * ρ 23 + i Ω p 1 ρ 32 + i Ω p 2 ρ 34 i Ω p 2 * ρ 43 + Γ 43 ρ 44 + Γ 23 ρ 22 ,
ρ 44 t = i Ω p 2 ρ 34 + i Ω p 2 * ρ 43 + i Ω c 2 * ρ 45 i Ω c 2 ρ 54 Γ 4 ρ 44 ,
ρ 55 t = i Ω c 2 * ρ 45 + i Ω c 2 ρ 54 + Γ 45 ρ 44 + Γ 25 ρ 22 ,
ρ 12 t = i δ c 1 ρ 12 i Ω c 1 * ( ρ 22 ρ 11 ) + i Ω p 1 * σ 13 Γ 2 + γ 12 2 ρ 12 ,
ρ 13 t = i ( δ c 1 δ p ) ρ 13 + i Ω p 1 ρ 12 + i Ω p 2 ρ 14 i Ω c 1 * ρ 23 γ 13 2 ρ 13 ,
ρ 14 t = i ( Δ + δ c 1 ) ρ 14 + i Ω p 2 * ρ 13 + i Ω c 2 * ρ 15 i Ω c 1 * ρ 24 Γ 4 + γ 14 2 ρ 14 ,
ρ 15 t = i ( Δ + δ c 1 δ c 2 ) ρ 15 + i Ω c 2 ρ 14 i Ω c 1 * ρ 25 γ 15 2 ρ 15 ,
ρ 23 t = i δ p ρ 23 i Ω c 1 ρ 13 + i Ω p 2 ρ 24 i Ω p 1 ( ρ 33 ρ 22 ) Γ 2 + γ 23 2 ρ 23 ,
ρ 24 t = i Δ ρ 24 i Ω c 1 ρ 14 + i Ω c 2 * ρ 25 + i Ω p 2 * ρ 23 i Ω p 1 ρ 34 Γ 2 + Γ 4 + γ 24 2 ρ 24 ,
ρ 25 t = i ( Δ δ c 2 ) ρ 25 i Ω c 1 ρ 15 i Ω p 1 ρ 35 + i Ω c 2 ρ 24 Γ 2 + γ 25 2 ρ 25 ,
ρ 34 t = i ( δ p + Δ ) ρ 34 i Ω p 1 * ρ 24 + i Ω c 2 * ρ 35 i Ω p 2 * ( ρ 44 ρ 33 ) Γ 4 + γ 34 2 ρ 34 ,
ρ 35 t = i ( δ p + Δ δ c 2 ) ρ 35 i Ω p 1 * ρ 25 i Ω p 2 * ρ 45 + i Ω c 2 ρ 34 γ 35 2 ρ 35 ,
ρ 45 t = i δ c 2 ρ 45 i Ω c 2 ( ρ 55 ρ 44 ) i Ω p 2 ρ 35 Γ 4 + γ 45 2 ρ 45 ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.