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

Controlling transverse shift of the reflected light via high refractive index with zero absorption

Open Access Open Access

Abstract

We present a theoretical investigation on controlling the transverse shift while most of the researches are on longitudinal Goos-Hänchen shift. A two-layer system is considered. The refractive index of the first layer is fixed. The second layer is an atomic system coupled by a strong laser field to realize the Λ-style electromagnetically induced transparency, and an additional microwave field drives the transition between the lower two levels to construct high refractive index with zero absorption. We use such phenomenon to modify the refractive index, and consequently the transverse shift in reflection. The properties of the atomic system and the transverse shift of reflected field are briefly studied. Our investigation shows that the shift can be tuned by the strength of the microwave field. And since the atomic system is quite sensitive to the phase of the light fields, through which the transverse shift can be manipulated effectively. More importantly, the absorption is limited due to the presence of the microwave field.

© 2017 Optical Society of America

1. Introduction

Reflection happens when the light impinges on the surface of two homogeneous isotropic media, and this phenomenon can be perfectly described by the Fresnel formulas [1]. For confined fields, the center of the reflected- light “spot” on the boundary surface between the two media is not coincident with that of the incident. There are two kinds of shifts. The one parallel with the incident plane is called the Goos-Hänchen shift (GHS) [2], and the other one perpendicular to the incident plane is named the Imbert-Fedorov shift [3], or transverse shift (TS). Interestingly, these two kinds of shift can be described in an unified theory [4–7]. The prediction of TS is based on the nonvanishing transverse energy flux of the evanescent field inside the media [8]. In fact, the TS can be found in reflection or refraction of an incident beam when it has a non-zero spin momentum (i.e., is not of linearly polarized) [9, 10]. The reason for such universality is that the TS is a corollary of the conservation of spin angular momentum and orbital angular momentum [11]. On the other hand, TS can be understood as a different version of spin Hall effect of light (SHEL) [11] which is a phenomenon about splitting of a linearly polarized incident beam into two circularly polarized light in reflection or refraction.

Recently, manipulation of GHS in the fixed configuration or device got a lot of attentions. One way to control GHS is to modify the refractive index of the media by applying a strong laser field [12, 13]. Then electromagnetically induced transparency [14] (EIT) seems to be a promising tool. Take the three-level Λ system as an example, EIT can be realized by applying a coupling field and a probe field driving respectively two adjacent transitions that share a common exited atomic level. The commonly used media for EIT are atomic gas [15, 16], and solid material, such as Pr3+-doped Y2SiO5 or Pr:YSO [17, 18], and nitrogen-vacancy (N-V) centers in diamond [19–21]. In the transparency window, the probe field experiences nearly zero absorption and steep positive dispersion [22]. Coherent control of GHS using EIT has been discussed in [23]. And most recently, M. S. Zubairy and his coworkers discussed controlling the GHS and TS in a cavity containing three-level atomic media via both pump and driving fields [24].

For the EIT system, in the middle of the transparency window (double-photon detuning δ = 0), the refractive index n = 1. For δ ≠ 0, the refractive index becomes larger, so does the absorption. The system is powerful for suppressing the absorption, but week for manipulating the refractive index. When the TS of the reflected field becomes our object of interest, we certainly wish that the absorption can be limited. To accomplish this, the high refractive index with zero absorption (HRIOA) [25–28] is a promising tool. One can introduce a microwave field (6.8 GHz for 87Rb atomic gas) into the system driving the two lower atomic levels. In such system, one can find both reduced and enhanced refractive index with zero absorption so that the reflected or refracted light beams do not suffer too much absorptive losses; Such reduced refractive index can be all-optically manipulated to get an enhanced difference in refraction coefficients for a high background refractive index so that a large enough transverse shift can be observed; Since the refractive index is tunable, and the absorption is limited, this phenomenon seems perfect for manipulating the TS.

In this article we present the research on controlling TS of the reflected field by HRIOA, which is realized by introducing a microwave field driving the lower atomic levels in the three-level Λ system. We adopt the formulas given in [9, 10] to present the analytical result of TS. The formulas are experimentally confirmed in [29]. The incident field is assumed to be the right-handed circularly polarized light. In Sec. 2, we present the model of three-level Λ system, and the equations describing the TS of the reflected field. Sec. 3 deals with manipulation of TS via the strength and the phase of the microwave field. We summarize our results and discussion in Sec. 4.

2. Atomic media and transverse shift of the reflected light

Let us consider a two-layer system composed by the media with the fixed permittivity ε1 and a layer of cold 87Rb atomic gas with the tunable permittivity ε2, as shown in Fig. 1(a). Both the GHS and the TS are marked out. The atomic media is modelled by the three-level Λ system. The schematic is presented in Fig. 1(b). For example, the three atomic levels could be |1〉 = |5S1/2, F = 1〉, |2〉 = |5S1/2, F = 2〉, and |3〉 = |5P1/2, F′= 2〉. When the Zeeman structure is considered, a static magnetic field should be applied to lift the degeneracy, In order to suppress the redundant transition [30]. The coupling field Ωc drives the transition |3〉 − |2〉, and a weak field Ωp probes the transition |3〉 − |1〉. The transition between the lower levels is driven by a microwave Ωd. The Rabi frequencies of the fields are defined as the complex numbers Ω˜c=Ωceiϕc=(32Ec/2)eiϕc, Ω˜d=Ωdeiϕd=(21Ed/2)eiϕd, Ω˜p=Ωpeiϕp=(31Ep/2)eiϕp. Where ℘ij is the matrix element of the dipole moment between levels |i〉 and |j〉. And Ex (x ∈ {c, d, p}) is the amplitude of corresponding field, ϕx denotes the phase of the field Ωx. The density equations that describes the atomic system are:

σ˙22=iΩd(σ12eiϕσ21eiϕ)+iΩc(σ32σ23)+Γ32σ33.
σ˙33=iΩp(σ13σ31)+iΩc(σ23σ32)(Γ31+Γ32)σ33.
σ˙23=iΩdeiϕσ13iΩpσ21iΩc(σ22σ33)+(iΔcγ)σ23.
σ˙13=iΩdeiϕσ23iΩcσ12+iΩp(σ222σ331)+(iΔpγ)σ13.
σ˙12=iΩdeiϕ(2σ22+σ331)+iΩpσ32iΩcσ13+(iδγm)σ12.

 figure: Fig. 1

Fig. 1 (a) Schematic of the transverse shift and the Goos-Hänchen shift of the reflected field on the surface of EIT media. (b) The three-level Λ system of 87Rb with a microwave field coupling the lower levels.

Download Full Size | PDF

Here ϕ = ϕd + ϕcϕp. The detunings of the fields are defined as Δc = ω32ωc, Δp = ω31ωp, δ = Δp − Δc with ωc (ωp) being the frequency of the coupling (probe) field, and ω32 and ω31 as the resonant frequencies of the corresponding optical transitions. Γij is the population decay rate between |i〉 and |j〉. We assume that the microwave Ωd resonates with the transition |2〉 − |1〉. γm is the coherence decay rate of the spin transition |2〉 − |1〉. For both optical transitions (|3〉 − |1〉 and |3〉 − |2〉), the coherence decay rates are assumed to have the same value denoted by γ. To the first order of the probe and the microwave fields, the density element σ31 can be written as

σ31(1)=σ13(1)*=ΩcΩdeiϕΩp(iγmδ)(iγmδ)(iγΔp)Ωc2.
The numerical solution of Eqs. (2) is given in Fig. 2. As we can see, when the refractive index (∝ Reσ31) reaches the maximum or minimum value, the absorption (∝ Imσ31) vanishes. This is the effect which has been referred to as HRIOA. The vertical magenta dashed lines indicate the probe detunings which satisfy Imσ31p) = 0. The points P1 and P2 are the intersection points of the vertical lines and line of Re σ31, and they can be used to represents the two HRIOA states in this article.

 figure: Fig. 2

Fig. 2 Reσ31 (red solid line) and Imσ31 (dark-blue dashed line) plotted as the function of Δp. The data is calculated from Eqs. (2) under the condition of σ˙ij=0. The parameters are Ωc = 3γ, Ωd = 0.5γ, Ωp = 0.01γ, γ = 2π × 6MHz, Δc = 0, γm = 0, ϕ = π/2, The horizontal magenta dashed line indicates σ31 = 0. And the vertical lines denote Δp = −2.9γ and Δp = 2.9γ at which Imσ31 = 0. P1 and P2 are the intersection points.

Download Full Size | PDF

Without the microwave Ωd, EIT is realized. The important feature of EIT is that, with a resonant coupling field Ωc, there are two absorption peaks located at Δp = ±Ωc, and Im σ31 approaches zero at Δp = 0 [14]. When the microwave Ωd is applied, the closed-loop transitions |1〉 → |2〉 → |3〉 → |1〉 and |1〉 → |3〉 → |2〉 → |1〉 may occur with the coherence of two ground states being perturbed to result in either constructive or destructive interference. Consequently, enhanced or suppressed EIT peaks can be obtained depending especially on the relative phase ϕ between the optical fields and the microwave field. From another point of view, for the proper values of the parameters, the enhancement of Ωp takes place as a consequence of the Raman process |1〉 → |2〉 → |3〉 → |1〉 while the probe absorption naturally exists in a nearby spectral region. At the “junction” of the attenuation and the amplification section, HRIOA can be realized. Under the assumption Ωc > γ > Ωd ≫ Ωp, and ϕ = π/2, the extrema of Reσ31 are ±Ωd/γ approximately, and located at Δp ≈ ∓Ωc. With the vanishing absorption and tunable refractive index, this effect is very promising for controlling the TS of the reflected light. The permittivity of the atomic media can be calculated by ε2=1+(312Nσ31)/(ε0Ωp), and the introduction of this relation and a comprehensive description of refractive index enhancement without absorption can be found in the famous textbook [31].

The TS of the reflected field can be calculated through the formulas given in [9, 10], and the general form is s = st + sa with

st=cotθk(1+|β|2+2Reβ)Imm+2RemImβ1+|βm|2
sa=cotθk1|β|21+|βm|2RemImDReD
Here θ denotes the incident angle. k = 2π/λ is the amplitude of wave vector. The parameter m describes the polarization of the incident field through the Jones vector (1,m)T/1+|m|2. For example, the Jones vector with m = −i corresponds to the right-handed circularly polarized light. D is a complex parameter characterizing the width and phase front curvature of the beam and relating to the distance propagation distance at which sa is measured, see [9] for detailed expressions. β = R/R, with
R=ε1cosθξε1cosθ+ξ.
R=ε2cosθε1ξε2cosθ+ε1ξ.

Where ξ = ε2ε1 sin2 θ. The shift sa is the angular transverse shift, it vanishes for the circularly polarized light. TS st given in Eq.(7) is what we are interested. The dimensionless transverse shift (s/λ) is given in Fig. 3. The Rabi frequencies are chosen as Ωc = 3γ, Ωd = 0.5γ, Ωp = 0.01γ. Then HRIOA happens at Δp = ±2.9γ ≈ ±Ωc. Due to the absence of absorption (Im ε2 = 0), β is a real number for partial reflection. It only becomes complex when total reflection occurs (ε2ε1 sin2 θ < 0). In Fig. 3, a red-solid (blue-dashed) curve corresponding to reduced (enhanced) refractive index with zero absorption is plotted for Δp = 2.9γp = −2.9γ) with ϕ = π/2. The maximal transverse shift is seen to occur at the critical angle of total reflection in both red-solid and blue-dashed curves because a largest extrinsic orbital angular momentum is required to satisfy the conservation law of total angular momentum when the transmitted field vanishes [32]. The maximal shift in the red-solid curve is higher than that in the blue-dashed curve, this is due to the fact that the contrast ϵ1/ϵ2 is larger in the case of reduced refractive index than that in the case of enhanced refractive index.

 figure: Fig. 3

Fig. 3 The dimensionless transverse shift s/λ with different probe detunings, Δp = 2.9γ (red solid line) and Δp = −2.9γ (dark-blue dashed line). The incident angles for the maximum shift (θ = 0.056π, and 0.212π rad) are marked by the magenta dashed line. ε1 = 2.22, and N312/ε0=0.05γ. Other parameters are identical with that in Fig. 2.

Download Full Size | PDF

3. Manipulating the transverse shift

In addition to realize HRIOA in the EIT media, The microwave Ωd provides effective ways to control the TS through the strength (Ωd) and the relative phase (ϕ).

3.1. Tuning the strength of microwave field

Fixing the intensity of the probe field and coupling field (Ωp = 0.01γ, Ωc = 3γ), with the previous assumption Δc = 0, Reσ31 depends on Ωd, Δp, and ϕ. Here we set ϕ = π/2, the values of Re σ31 for P1 and P2 are given in Fig 4(a) as the function of Ωd. For weaker microwave field (Ωd < 0.2γ), Re σ31 ∝ Ωd which is consistent with the properties of the HRIOA we mentioned in the previous section. The values of Δp for P1 and P2 are presented in Fig. 4(b) as a function of Ωd as well. As we can see that Δp approximately remains constant as Ωc for P2, and −Ωc for P1. This is due to the feature of EIT.

 figure: Fig. 4

Fig. 4 (a) The value of Reσ31 for P1 (dark-blue dashed line) and P2 (red solid line) plotted as the function of Ωd. The relations between Δp and Ωd at points P1 and P2 are given in (b). And (c) shows the largest TS of the reflected light when HRIOA is achieved. The incident angle is the critical angle for total reflection. Other parameters are Ωc = 3γ, Ωp = 0.01γ, γ = 2π × 6MHz, ϕ = π/2, ε1 = 2.22, and N312/ε0=0.05γ.

Download Full Size | PDF

The relation given in Fig. 4(b) guarantees that the atomic system stays in HRIOA states. Without absorption or amplification of the incident light, ε2 remains a real number. In this subsection, we assume that the system is always in HRIOA states. The maximal TS of the reflected field corresponding to P1 and P2 are calculated, and shown in Fig. 4(c). Defining n=ε2/ε1, and the maximum of the shift for partial reflection can be easily calculated as

spm=λr(4r2)3/22π(8+2(n23)r2+r4).
The corresponding incident angle is θpm = arccos (r/2), with r=n1n4+10n2+1n21. For total reflection, the maximum value of the shift is
stm=λ1n2πn,
which happens at the critical angle of total reflection θtm = arcsin (n). Normally, stm > spm. Here we would like to briefly discuss the underlying physics of the above equation. The transverse shift relates to the conservation of the total angular momentum along the normal direction (zaxis) of the interface. For total reflection, we have SzrSzi=stmksinθ. The left-hand side of the equation is change of “z-component” of the spin between the incident and reflected fields, and the right-hand side is the change of the extrinsic orbital momentum. At the critical angle, θ = θc, we have cosθc=1n2, and SzrSzi=2cosθc=stmksinθc. Then the expression of stm can be obtained, and it is the Eq. (12). One can see that a small relative refractive index n leads to more significantly changed spin momentum of the field after reflection, and this causes a larger transverse shift based on the conservation of the total angular momentum. Detailed discussion on this topic can be found in [32]. The parameters that we choose allows total reflection for both P1 and P2, therefore, the incident angle in Fig. 4(c) is θtm and the data in Fig. 4(c) is connected with that in Fig. 4(a) through the Eq. (12). Due to this property, the maximum shift with Δp ≈ Ωc (red solid line) is larger than that with Δp ≈ −Ωc (blue dashed line). And more importantly the former situation offers the more flexible control of the TS through adjusting Ωc.

3.2. Tuning the phase of microwave field

The system of HRIOA is very sensitive to the relative phase ϕ [27]. Taking advantage of this feature, the TS can be effectively manipulated. Fig. 5(a) shows the real and imaginary parts of density matrix element σ31 with ϕ = −π/2. Roughly speaking, σ31 (ϕ = −π/2) ≈ −σ31 (ϕ = −π/2) comparing the results shown in Fig. 2 where ϕ = π/2. Set Ωc = 3γ, Ωd = 0.5γ, Ωp = 0.01γ, and Δp = 2.9γ, the TS with ϕ = π/2 and ϕ = −π/2 are presented in Fig. 5(b). The result is much similar with Fig. 3, however it reveals how effectively the TS can be manipulated by the relative phase ϕ.

 figure: Fig. 5

Fig. 5 (a) Re σ31 (red solid lines) and Im σ31 (dark-blue dashed lines) versus the probe detuning Δp with ϕ = −π/2, (b) Transverse shifts with ϕ = π/2 (dark-blue dashed line) and ϕ = π/2 plotted as the incident angle. Δp = 2.9γ. The other parameters are identical with Fig. 2.

Download Full Size | PDF

Before considering the manipulation of TS continuously by tuning ϕ, we first focus on the absorption and amplification of the system. With certain Ωc, Ωd, and set Δp = −Ωc, we can treat σ31 (which is from the solution of Eq.(2) with σ˙ij=0) as the function of ϕ. And by solving the equation Imσ31 (ϕ) = 0, we get ϕ = ±ϕ0, with ϕ0 = arccos (Ωpd) under the condition of Ωc > γ > Ωd ≫ Ωp. The process of obtaining the above solution is trivial and complicated, therefore we shall not present it here. Set Ωd = 0.5γ and Ωp = 0.01γ, we have ϕ0 = ±0.49π. In Fig. 6(a), we present the numerical solution of σ31 obtained from Eq. (2). The imaginary part of σ31 corresponds to the absorption or amplification of the probe field. Here the positive Im σ31 means absorption, while the negative one means amplification. In the previous discussion, the parameter β becomes a complex value when total reflection happens. Here when the non-zero Im σ31 is considered, β becomes complex as well since it relate to ε2 (which is determined by σ31). In order to obtain β, we have to calculate the term ξ in Eqs. (2). This value actually contributes to the wave vector of the probe light in the EIT media. Set kn as the projection of the wave vector onto the normal direction of the boundary surface, then kn=ξk=|ξ|keiθξ/2. As we already knew that, when light is totally reflected by the gain media, it get amplified [33], therefore θξ should be in the range −π/2 < θξ ≤ 3π/2 [34]. For total reflection (θ>arcsin(Reε2)/ε1) the means Re kn < 0 (evanescent wave) and Im kn > 0 (gain). Fig. 6(b) shows the reflection coefficient of the circular light (m = −i), R=12(R2+R2)1/2. As we can see that R > 1 when the total reflection takes place with the negative Im σ31.

 figure: Fig. 6

Fig. 6 (a) Re σ31 (red solid line) and Im σ31 (dark-blue dashed line) plotted as the function of ϕ. (b) The reflection coefficient of the circularly polarized light (m = −i). The parameters are Ωc = 3γ, Ωp = 0.01γ, Ωd = 0.5Ωc, Δp = −Ωc, γ = 2π × 6MHz, ε1 = 2.22.

Download Full Size | PDF

We stimulate the TS of the reflected light for different relative phases ϕ and the incident angles θ. The result is given in Fig. 7(a). The detailed behavior of TS in the yellow rectangle is presented in Fig. 7(b). There are two peaks of TS in the ϕθ plane, corresponding to the total reflection with HRIOA. The TS is a result of conservation law for normal-direction component of total angular momentum [10]. When tuning ϕ with the other parameters fixed, the EIT medium might become active. The absorption or amplification can break this conservation, and result in this two-peaks structure.

 figure: Fig. 7

Fig. 7 (a) The TS of the reflected light versus the relative phase ϕ and the incident angle θ. The detailed behavior of the TS in the yellow rectangle is presented in (b). N312/ε0=0.05γ. The other parameters are identical with that in Fig. 6.

Download Full Size | PDF

It is of interest to discuss what will happen when the medium is a hot atomic vapor. In the presence of Doppler broadening, EIT windows become much shallower to exhibit more residual absorption. Fortunately, a so-called Doppler-free technique [35] can cancel the difference of Doppler shifts on different transitions, yielding roughly same EIT spectra as in cold atomic gases. Similarly, we expect to attain roughly same results on HRIOA when the microwave Ωd is applied. EIT can also be realized in the solid materials, such as Pr:YSO, and N-V centers in diamond. The transition frequency between the lower levels are 10.2 MHz for Pr:YSO [17, 18] which means a radio-frequency field should be used to couple those two levels. and for negatively charged N-V centers, the microwave of frequency 2.88 GHz [36] is reasonable to realize HRIOA.

4. Conclusion

In this article we investigated the transverse shift (TS) on the surface of EIT media. A microwave field is employed to couple the two lower atomic levels, in order to construct high refractive index with zero absorption (HRIOA). Using this effect, we can control the refractive index, and furthermore, the TS of the reflected field. We briefly investigated the properties of HRIOA. When all the intensities of the fields are fixed, normally Ωc > γ > Ωd ≫ Ωp, and the coupling field and the microwave field drive the corresponding transition resonantly. There are two value of Δp that leads to HRIOA. Consequently, there exists two type of the TS. The TS of the reflected field can be controlled by the strength and the phase of the microwave field. We maintain the system in the HRIOA state, and change the strength of the microwave field to investigate the maximum shift that can be obtained. The result is given in Fig. 4. For the phase of the microwave, the EIT media can become active (leave the state of HIROA) as the phase changed continuously meanwhile the other parameters fixed. Two peaks of TS are found in the ϕθ plane, correspond to the two states of HRIOA of the EIT system. And this means that the TS can respond to the phase of microwave fields sensitively. Our investigation is done for a two-layer system composed by a medium with a fixed refractive index and a medium with a controllable refractive index. To implement a relevant experimental research, however, more details need to be taken into account like transverse shifts at the interface to the containing vessel and polarization changes under an oblique incident angle. With these considerations, the polarization giving a maximum shift at the critical incidence may be not exactly circular [37] as we show here.

Funding

National Natural Science Foundation of China (Grants No. 61378094, 11534002, 11404336, 11374126).

Acknowledgments

We thank Professor Ren-Gang Wan for helpful discussions.

References and links

1. M. A. X. Born and E. Wolf, Principles of Optics (Seventh (Expanded) Edition) (Cambridge University, 1999).

2. F. Goos and H. Hänchen, “Ein neuer und fundamentaler versuch zur totalreflexion,” Ann. Phys. 436, 333–346 (1947). [CrossRef]  

3. F. I. Fedorov, “On the theory of total internal reflection,” Dokl. Akad. Nauk SSSR 105, 465 (1955).

4. C.-F. Li, “Unified theory for Goos-Hänchen and Imbert-Fedorov effects,” Phys. Rev. A 76, 013811 (2007). [CrossRef]  

5. A. Aiello, “Goos-Hänchen and Imbert-Fedorov shifts: a novel perspective,” New J. Phys. 14, 013058 (2012). [CrossRef]  

6. A. Aiello, M. Merano, and J. P. Woerdman, “Duality between spatial and angular shift in optical reflection,” Phys Rev. A 80, 061801 (2009). [CrossRef]  

7. T. Paul, C. Rockstuhl, C. Menzel, and F. Lederer, “Resonant Goos-Hänchen and Imbert-Fedorov shifts at photonic crystal slabs,” Phys. Rev. A 77, 053802 (2008). [CrossRef]  

8. C. Imbert, “Calculation and experimental proof of the transverse shift induced by total internal reflection of a circularly polarized light beam,” Phys. Rev. D 5, 787–796 (1972). [CrossRef]  

9. K. Y. Bliokh and Y. P. Bliokh, “Polarization, transverse shifts, and angular momentum conservation laws in partial reflection and refraction of an electromagnetic wave packet,” Phys. Rev. E 75, 066609 (2007). [CrossRef]  

10. K. Y. Bliokh and Y. P. Bliokh, “Conservation of angular momentum, transverse shift, and spin Hall effect in reflection and refraction of an electromagnetic wave packet,” Phys. Rev. Lett. 96, 073903 (2006). [CrossRef]   [PubMed]  

11. M. Onoda, S. Murakami, and N. Nagaosa, “Hall effect of light,” Phys. Rev. Lett. 93, 083901 (2004). [CrossRef]   [PubMed]  

12. L.-G. Wang, M. Ikram, and M. S. Zubairy, “Control of the Goos-Hänchen shift of a light beam via a coherent driving field,” Phys. Rev. A 77, 023811 (2008). [CrossRef]  

13. Ziauddin and S. Qamar, “Control of the Goos-Hänchen shift using a duplicated two-level atomic medium,” Phys. Rev. A 85, 055804 (2012). [CrossRef]  

14. S. E. Harris, “Electromagnetically induced transparency,” Phys. Today 50, 36 (1997). [CrossRef]  

15. D. F. Phillips, A. Fleischhauer, A. Mair, R. L. Walsworth, and M. D. Lukin, “Storage of light in atomic vapor,” Phys. Rev. Lett. 86, 783–786 (2001). [CrossRef]   [PubMed]  

16. C. Liu, Z. Dutton, C. H. Behroozi, and L. V. Hau, “Observation of coherent optical information storage in an atomic medium using halted light pulses,” Nature 409, 490–493 (2001). [CrossRef]   [PubMed]  

17. B. S. Ham, P. R. Hemmer, and M. S. Shahriar, “Efficient electromagnetically induced transparency in a rare-earth doped crystal,” Opt. Commun. 144, 227–230 (1997). [CrossRef]  

18. B. S. Ham, M. S. Shahriar, M. K. Kim, and P. R. Hemmer, “Frequency-selective time-domain optical data storage by electromagnetically induced transparency in a rare-earth-doped solid,” Opt. Lett. 22, 1849–1851 (1997). [CrossRef]  

19. C. Wei and N. B. Manson, “Observation of the dynamic stark effect on electromagnetically induced transparency,” Phys. Rev. A 60, 2540–2546 (1999). [CrossRef]  

20. C. Wei and N. B. Manson, “Observation of electromagnetically induced transparency within an electron spin resonance transition,” J. Opt. B: Quantum Semiclass. Opt. 1, 464 (1999). [CrossRef]  

21. P. R. Hemmer, A. V. Turukhin, M. S. Shahriar, and J. A. Musser, “Raman-excited spin coherences in nitrogen-vacancy color centers in diamond,” Opt. Lett. 26, 361–363 (2001). [CrossRef]  

22. M. Fleischhauer, A. Imamoglu, and J. P. Marangos, “Electromagnetically induced transparency: Optics in coherent media,” Rev. Mod. Phys. 77, 633–673 (2005). [CrossRef]  

23. Ziauddin, S. Qamar, and M. S. Zubairy, “Coherent control of the Goos-Hänchen shift,” Phys. Rev. A 81, 023821 (2010). [CrossRef]  

24. S. Asiri, J. Xu, M. Al-Amri, and M. S. Zubairy, “Controlling the Goos-Hänchen and Imbert-Fedorov shifts via pump and driving fields,” Phys. Rev. A 93, 013821 (2016). [CrossRef]  

25. S. E. Harris, “Refractive-index control with strong fields,” Opt. Lett. 19, 2018–2020 (1994). [CrossRef]   [PubMed]  

26. A. S. Zibrov, M. D. Lukin, L. Hollberg, D. E. Nikonov, M. O. Scully, H. G. Robinson, and V. L. Velichansky, “Experimental demonstration of enhanced index of refraction via quantum coherence in Rb,” Phys. Rev. Lett. 76, 3935–3938 (1996). [CrossRef]   [PubMed]  

27. W.-H. Xu, J.-H. Wu, and J.-Y. Gao, “Effects of decay-induced coherence and microwave-induced coherence on the index of refraction in a three-level Λ-type atomic system,” Laser Phys. Lett. 1, 176–183 (2004). [CrossRef]  

28. H.-F. Zhang, J.-H. Wu, X.-M. Su, and J.-Y. Gao, “Quantum-interference effects on the index of refraction in an Er3+-doped yttrium aluminum garnet crystal,” Phys. Rev. A 66, 053816 (2002). [CrossRef]  

29. O. Hosten and P. Kwiat, “Observation of the spin Hall effect of light via weak measurements,” Science 319, 787–790 (2008). [CrossRef]   [PubMed]  

30. X.-G. Wei, J.-H. Wu, G.-X. Sun, Z. Shao, Z.-H. Kang, Y. Jiang, and J.-Y. Gao, “Splitting of an electromagnetically induced transparency window of rubidium atoms in a static magnetic field,” Phys. Rev. A 72, 023806 (2005). [CrossRef]  

31. M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University, 1997).

32. K. Y. Bliokh and A. Aiello, “Goos-Hänchen and Imbert-Fedorov beam shifts: an overview,” J. Opt. 15, 014001 (2013). [CrossRef]  

33. C. Koester, “Laser action by enhanced total internal reflection,” IEEE J. of Quantum Electron. 2, 580–584 (1966). [CrossRef]  

34. K. J. Willis, J. B. Schneider, and S. C. Hagness, “Amplified total internal reflection: theory, analysis, and demonstration of existence via fdtd,” Opt. Express 16, 1903–1914 (2008). [CrossRef]   [PubMed]  

35. Y.-q. Li and M. Xiao, “Electromagnetically induced transparency in a three-level Λ-type system in rubidium atoms,” Phys. Rev. A 51, R2703–R2706 (1995). [CrossRef]  

36. C. Santori, D. Fattal, S. M. Spillane, M. Fiorentino, R. G. Beausoleil, A. D. Greentree, P. Olivero, M. Draganski, J. R. Rabeau, P. Reichart, B. C. Gibson, S. Rubanov, D. N. Jamieson, and S. Prawer, “Coherent population trapping in diamond n-v centers at zero magnetic field,” Opt. Express 14, 7986–7994 (2006). [CrossRef]   [PubMed]  

37. J. B. Götte, W. Löffler, and M. R. Dennis, “Eigenpolarizations for giant transverse optical beam shifts,” Phys. Rev. Lett. 112, 233901 (2014). [CrossRef]   [PubMed]  

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 (a) Schematic of the transverse shift and the Goos-Hänchen shift of the reflected field on the surface of EIT media. (b) The three-level Λ system of 87Rb with a microwave field coupling the lower levels.
Fig. 2
Fig. 2 Reσ31 (red solid line) and Imσ31 (dark-blue dashed line) plotted as the function of Δp. The data is calculated from Eqs. (2) under the condition of σ ˙ i j = 0. The parameters are Ωc = 3γ, Ωd = 0.5γ, Ωp = 0.01γ, γ = 2π × 6MHz, Δc = 0, γm = 0, ϕ = π/2, The horizontal magenta dashed line indicates σ31 = 0. And the vertical lines denote Δp = −2.9γ and Δp = 2.9γ at which Imσ31 = 0. P1 and P2 are the intersection points.
Fig. 3
Fig. 3 The dimensionless transverse shift s/λ with different probe detunings, Δp = 2.9γ (red solid line) and Δp = −2.9γ (dark-blue dashed line). The incident angles for the maximum shift (θ = 0.056π, and 0.212π rad) are marked by the magenta dashed line. ε1 = 2.22, and N 31 2 / ε 0 = 0.05 γ. Other parameters are identical with that in Fig. 2.
Fig. 4
Fig. 4 (a) The value of Reσ31 for P1 (dark-blue dashed line) and P2 (red solid line) plotted as the function of Ωd. The relations between Δp and Ωd at points P1 and P2 are given in (b). And (c) shows the largest TS of the reflected light when HRIOA is achieved. The incident angle is the critical angle for total reflection. Other parameters are Ωc = 3γ, Ωp = 0.01γ, γ = 2π × 6MHz, ϕ = π/2, ε1 = 2.22, and N 31 2 / ε 0 = 0.05 γ.
Fig. 5
Fig. 5 (a) Re σ31 (red solid lines) and Im σ31 (dark-blue dashed lines) versus the probe detuning Δp with ϕ = −π/2, (b) Transverse shifts with ϕ = π/2 (dark-blue dashed line) and ϕ = π/2 plotted as the incident angle. Δp = 2.9γ. The other parameters are identical with Fig. 2.
Fig. 6
Fig. 6 (a) Re σ31 (red solid line) and Im σ31 (dark-blue dashed line) plotted as the function of ϕ. (b) The reflection coefficient of the circularly polarized light (m = −i). The parameters are Ωc = 3γ, Ωp = 0.01γ, Ωd = 0.5Ωc, Δp = −Ωc, γ = 2π × 6MHz, ε1 = 2.22.
Fig. 7
Fig. 7 (a) The TS of the reflected light versus the relative phase ϕ and the incident angle θ. The detailed behavior of the TS in the yellow rectangle is presented in (b). N 31 2 / ε 0 = 0.05 γ. The other parameters are identical with that in Fig. 6.

Equations (12)

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

σ ˙ 22 = i Ω d ( σ 12 e i ϕ σ 21 e i ϕ ) + i Ω c ( σ 32 σ 23 ) + Γ 32 σ 33 .
σ ˙ 33 = i Ω p ( σ 13 σ 31 ) + i Ω c ( σ 23 σ 32 ) ( Γ 31 + Γ 32 ) σ 33 .
σ ˙ 23 = i Ω d e i ϕ σ 13 i Ω p σ 21 i Ω c ( σ 22 σ 33 ) + ( i Δ c γ ) σ 23 .
σ ˙ 13 = i Ω d e i ϕ σ 23 i Ω c σ 12 + i Ω p ( σ 22 2 σ 33 1 ) + ( i Δ p γ ) σ 13 .
σ ˙ 12 = i Ω d e i ϕ ( 2 σ 22 + σ 33 1 ) + i Ω p σ 32 i Ω c σ 13 + ( i δ γ m ) σ 12 .
σ 31 ( 1 ) = σ 13 ( 1 ) * = Ω c Ω d e i ϕ Ω p ( i γ m δ ) ( i γ m δ ) ( i γ Δ p ) Ω c 2 .
s t = cot θ k ( 1 + | β | 2 + 2 R e β ) I m m + 2 R e m I m β 1 + | β m | 2
s a = cot θ k 1 | β | 2 1 + | β m | 2 R e m I m D R e D
R = ε 1 cos θ ξ ε 1 cos θ + ξ .
R = ε 2 cos θ ε 1 ξ ε 2 cos θ + ε 1 ξ .
s p m = λ r ( 4 r 2 ) 3 / 2 2 π ( 8 + 2 ( n 2 3 ) r 2 + r 4 ) .
s t m = λ 1 n 2 π n ,
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.