Abstract
A coherently prepared asymmetric double semiconductor quantum well (QW) is proposed to realize parity-time (PT) symmetry. By appropriately tuning the laser fields and the pertinent QW parameters, PT-symmetric optical potentials are obtained by three different methods. Such a coherent QW system is reconfigurable and controllable, and it can generate new approaches of theoretically and experimentally studying PT-symmetric phenomena.
Similar content being viewed by others
Introduction
The non-Hermitian parity-time (PT)-symmetric Hamiltonians, which were firstly proposed by Bender and Boettcher in 1998, have attracted great attention1. Because of the isomorphism between quantum Schrödinger equations and paraxial-wave equations2, optical system becomes an ideal bed for experimentally studying PT symmetry. By balancing gain and loss, optical PT symmetry has been realized in different coupled structures, such as waveguides3,4, lattices5,6, micro-cavities7,8, and can be used in the fields of unidirectional propagating9,10, perfect absorbers11,12, photon lasers13,14, phonon lasers15 and sensors16. In addition, several interesting phenomena, such as optical solitons17, Bloch oscillations18 and topological insulators19 have also been investigated in optical PT-symmetric systems. However, most of the work, particularly in experiment, was based on solid-state materials. Once the structures are fabricated, the properties such as the threshold of the system are unable to be changed.
The susceptibility of host semiconductor quantum wells (QWs) can be controlled via electrical field20,21,22, carrier density23 and laser field24. The optical response of QWs is significantly enhanced when the frequency of light field is near resonance with the intersubband transition which has large dipole matrix element25. As a result, one can obtain a dramatic change in the complex dielectric constant20,21,22,23,24,25. Combined with electromagnetically induced transparency (EIT)26,27, both the refractive index and the absorptive coefficient of QWs can be effectively manipulated via atomic coherence and quantum interference.
In this work, we propose to use a coherently prepared asymmetric double semiconductor QWs to obtain PT symmetry. Such QW systems have been proved to have the possibility to realize quantum coherence and interference28,29,30. We demonstrate that PT-symmetric optical potentials, such as coupled optical waveguides, one-dimensional (1D) and two-dimensional (2D) PT-symmetric optical lattices can be realized by different methods. Besides, it is possible to control the PT-symmetric properties by tuning the laser fields and the QW parameters. The realization of PT symmetry in coherent QW systems by laser fields have several advantages. Firstly, compared with solid-state materials with micro-structures or nano-structures, PT-symmetric properties in this system can be established by different methods, and can be effectively controlled by various parameters. Secondly, PT symmetry has been constructed theoretically31,32,33,34 and experimentally35 in different atomic systems. Compared with these atomic systems, semiconductor QW systems have designable and flexible of energy levels, and they are easy to be integrated and stable for practical application. Thirdly, in such systems large nonlinearity can be realized assisted by EIT36,37,38, which makes it possible to observe traveling effects of lights in non-Hermitian nonlinear optical systems39,40,41.
Models and Equations
We consider asymmetric double semiconductor QWs42, which consist of a 51-monolayer (145 Å)-thick wide well (WW) and 35-monolayer (100 Å)-thick narrow well (NW). Between them there is a thickness of a 9-monolayer (25-Å)-thick Al0.2Ga0.8As barrier, as shown in Fig. 1(a). There are ten pairs of QWs (each pair consists of one WW, one NW, and one thick barrier), which are isolated from each other by 200-Å-wide Al0.2Ga0.8As buffer layers. All these pairs are sandwiched between nominally undoped 3500-Å-thick Al0.2Ga0.8As layers.
Such asymmetric double semiconductor QWs can be treated as a four-level N configuration28,29,30, as shown in Fig. 1(b). Here, levels |1〉 and |2〉 are localized hole states in a valence band, while levels |3〉 and |4〉 are delocalized bonding and antibonding states in a conduction band, arising from the tunneling effect between the WW and NW via the thin barrier, respectively. The probe field Ep with frequency ωp probes the transition \(|1\rangle \leftrightarrow |3\rangle \), while the coupling field Ec with frequency ωc and the pump field Ed with frequency ωd act on transitions \(|2\rangle \leftrightarrow |3\rangle \) and \(|1\rangle \leftrightarrow |4\rangle \) respectively. The Rabi frequency of the probe, coupling, and pump fields are Ωp = μ13Ep/2h, Ωc = μ23Ec/2h, and Ωd = μ14Ed/2h, respectively, where µij is the associated dipole transition matrix element, and the detuning of the probe, coupling, and pump fields are Δp = ωp − ω31, Δc = ωc − ω32, and Δd = ωd − ω42, respectively, where ωij is the transition frequency between levels |i〉 and |j〉.
Under the condition of low QW carrier intensity, many-body effects are attributed to electron–electron interactions can be neglected43. In the interaction picture and under the rotating wave approximation, the Hamiltonian of the QW system can be written as (h = 1)
Here, H.c. is the Hamiltonian complex conjugate.
The equation of motion for the density matrix of the system under the relaxation process is
where γ is the dissipation matrix. Substituting Eq. (1) into Eq. (2), the density matrix for each element can be obtained:
Here ρ11 + ρ22 + ρ33 + ρ44 = 1 and \({\rho }_{ij}={\rho }_{ji}^{\ast }\). Γij is the natural decay rate between levels |i〉 and |j〉. We assume that the decay rates from levels |3〉 and |4〉 in the conduction band to levels |1〉 and |2〉 in the valence band are identical, there are, Γ31 = Γ32 and Γ41 = Γ42. There is also no decay in the valence band or conduction band, so we conclude that Γ21 = Γ43 = 0. \({{\rm{\Gamma }}}_{i}={\sum }_{j}^{i-1}{{\rm{\Gamma }}}_{ij}\) denotes the total decay rate of level |i〉. We define \({\tilde{\gamma }}_{12}\) = \({\gamma }_{12}+i({{\rm{\Delta }}}_{p}-{{\rm{\Delta }}}_{c})\), \({\tilde{\gamma }}_{13}\) = \({\gamma }_{13}+i{{\rm{\Delta }}}_{p}\), \({\tilde{\gamma }}_{14}\) = \({\gamma }_{14}+i{{\rm{\Delta }}}_{d}\), \({\tilde{\gamma }}_{23}\) = \({\gamma }_{23}+i{{\rm{\Delta }}}_{c}\), \({\tilde{\gamma }}_{24}\) = \({\gamma }_{24}+i({{\rm{\Delta }}}_{c}+{{\rm{\Delta }}}_{d}-{{\rm{\Delta }}}_{p})\), and \({\tilde{\gamma }}_{34}\) = \({\gamma }_{34}+i({{\rm{\Delta }}}_{d}-{{\rm{\Delta }}}_{p})\), where γij = (Γi + Γj)/2.
In such QW systems, \({{\rm{\Gamma }}}_{3}={{\rm{\Gamma }}}_{3l}+{{\rm{\Gamma }}}_{3}^{{\rm{dph}}}\) and \({{\rm{\Gamma }}}_{4}={{\rm{\Gamma }}}_{4l}+{{\rm{\Gamma }}}_{4}^{{\rm{dph}}}\), where Γ3l and Γ4l are the population decay rates of subbands |3〉 and |4〉 respectively, resulting from longitudinal-optical phonon emission events at low temperature, and \({{\rm{\Gamma }}}_{3}^{{\rm{dph}}}\) and \({{\rm{\Gamma }}}_{4}^{{\rm{dph}}}\) are the dephasing decay rates of quantum coherence due to electron–electron scattering, phonon scattering processes; and elastic interface roughness. Based on the previous studies, we assume that Γ3l can be equal to Γ4l, and \({{\rm{\Gamma }}}_{3}^{{\rm{dph}}}\) can be equal to \({{\rm{\Gamma }}}_{4}^{{\rm{dph}}}\). Therefore, Γ31 = Γ32 = Γ41 = Γ42 = Γ can be presented. In addition, we assume there is no interference or dephasing between levels |3〉 and |4〉, which can be realized based on the appropriate reduction of the temperature44.
The susceptibility of the QW medium can be obtained through the expression \(\chi \) = \(N{\mu }_{13}/{\varepsilon }_{0}{E}_{p}\cdot {\rho }_{31}\) = \(N{\mu }_{13}^{2}/2{\varepsilon }_{0}{{\rm{h}}{\rm{\Omega }}}_{p}\cdot {\rho }_{31}\), where N is the electron density of the QWs, and ρ31 can be obtained by Eq. (3). χ = χR + iχI, where χR describes the dispersion properties of the probe field, while χI describes the absorption properties of the probe field with χI > 0 (χI < 0) indicating loss (gain).
The refractive index of the QW medium can be written as \(n=\sqrt{{\varepsilon }_{c}+\chi }\). Here \({\varepsilon }_{c}={n}_{c}^{2}\) corresponding to the complex dielectric constant of the host QW medium, and nc is its refractive index when light is far detuned from the resonance. χ denotes the change in the susceptibility22,23, which results from the coherent contract via the coupling and pump fields near the resonance. Generally, χ = εc, then we have \(n\) ≈ \(\sqrt{{\varepsilon }_{c}}+\chi /22\sqrt{{\varepsilon }_{c}}\) = \({n}_{c}+\chi /2{n}_{c}\) = \({n}_{c}+{\chi }_{R}/2{n}_{c}+i{\chi }_{I}/2{n}_{c}\). We define the real and imaginary parts of the refractive index as nR and nI, with nc being the background index of the system. Thus, n = nc + nR + inI, where nR = χR/2n0 and nI = χI/2n0. To achieve PT symmetry, the condition of nR(r) = nR(−r) and nI(r) = −nI(−r) must be satisfied. And for simplicity, we use the unit \(N{\mu }_{13}^{2}/4{\varepsilon }_{0}{\rm{h}}{n}_{0}\) in the following calculations.
Results
PT-symmetric optical waveguides
In order to realize PT-symmetric optical waveguides, we use a pair of coupling laser beams to form two coupled waveguides. Such pair of beams propagate in QWs along y direction. Then, a pump and a probe laser beams with wider laser dimension propagate in the same direction as those of the coupling beams. The schematic diagram is shown in Fig. 2. The pair of coupling laser beams have an identical Gaussian intensity profile so that the total spatial intensity distribution of them varying in x direction is
where A and a are the constant and the half separation between the two waveguides respectively. \(2\sqrt{2\,\mathrm{ln}\,2}\sigma \) can describe the full width at half maximum (FWHM) of the waveguides. By choosing two different detuning of the coupling fields, gain can be introduced to one waveguide and loss can be introduced to the other, even though other parameters are identical. In such consideration, the refractive index in each waveguide spatially varies only with the intensity of the coupling.
To realize gain and absorption in two waveguides simultaneously, different detuning of the coupling fields need to be found. For sake of it, we calculate the real (dispersion) χR and imaginary (gain or absorption) χI parts of the susceptibility as a function of Δc, and show the results in Fig. 3(a,b), respectively. The intensity of the coupling field in each waveguide is corresponding to Gaussian profile, so χI needs to get larger with increasing coupling intensity. For this reason, we only consider a negative value of the coupling detuning. It can be seen from the figures that for different value of Ωc, χR reaches to the maximum value, and χI is close to zero at Δc = −2.325Γ. At the vicinity of the zero point, gain is obtained on the left side, and absorption is abstained on the right side. This property called refractive index enhancement with vanishing absorption, has been demonstrated in the early papers45,46,47,48,49.
(a) Real part χR and (b) imaginary part χI of the susceptibility as a function of coupling detuning Δc for different coupling Rabi frequency, Ωc = 0.4 meV (black solid line), Ωc = 0.3 meV (blue dashed line), and Ωc = 0.2 meV (red dotted line). (c) Real part χR and (d) imaginary part χI of the susceptibility as a function of coupling Rabi frequency Ωc for different coupling detuning, Δc = −2.278 meV (blue solid line) and Δc = −2.360 meV (red dashed line). The other parameters are \({{\rm{\Gamma }}}_{3}^{{\rm{dph}}}={{\rm{\Gamma }}}_{4}^{{\rm{dph}}}=2.58{\rm{meV}}\), Γ3l = Γ4l = 2.07 meV, Γ31 = Γ32 = Γ41 = Γ42 = Γ = 2.325 meV, Ωp = 0.01Γ, Ωd = 2Γ, Δp = Δd = 0.
In order to realize gain and loss in two waveguides simultaneously by using two different coupling detuning and maintain other parameters identical, the relation between the susceptibility and the coupling intensity needs to be established. Therefore, we choose Δc = −2.360Γ and Δc = −2.278Γ in two waveguides, and draw χR and χI as a function of Ωc shown in Fig. 3(c,d). It can be seen that with selected coupling detuning, the curves of χR are identical, and the curves of χI are matched with slight difference, respectively. It should be noted that, in Fig. 3(d) we flipped the curve of the gain one by multiplying a minus sign to compare the two curves of gain and loss directly.
Based on the above analysis, firstly, we demonstrate the possibility of realizing spatial modulation of the refractive index. Making use of Eq. (4), and choosing the FWHM (\(2\sqrt{2\,\mathrm{ln}\,2}\sigma \)) as 7 μm and the separation between the two waveguides as 20 μm, we plot the real nR and the imaginary nI parts of the refractive index as a function of x shown in Fig. 4(a–c). It can be seen from the figures that the real part of the refractive index nR is an even function of x, while that of imaginary part nI is an odd function of x. By changing the electron intensity of QW, The absolute values of nR and nI can be modified with equal scale simultaneously with the changes of the electron intensity of QW.
Real part nR and imaginary part nI of the refractive index as a function of position x for different cases, (a) Δc = −2.350 meV (loss waveguide) and Δc = −2.300 meV (gain waveguide), representing the condition of below threshold, (b) Δc = −2.386 meV (loss waveguide) and Δc = −2.266 meV (gain waveguide), representing the condition of above threshold, (c) Δc = −2.325 meV (both waveguides), representing the condition of non-PT symmetry, respectively. (d–f) are filed mode of the probe laser beam according to (a–c). (g–i) are propagation properties of the probe laser beam according to (a–c). The other parameters are the same as in Fig. 3.
By controlling the ratio of the real and imaginary parts of the refractive index, we can adjust the condition of the system below or above the PT-symmetric threshold. It can be seen from Fig. 3(a,b) that χR changes less than χI at the vicinity of zero point (Δc = −2.325Γ). Therefore, we can alter the coupling detuning in the waveguides to control the ratio. For instance, with different values of the coupling detuning, the ratio of the real and imaginary parts is 12 and 80, and the system can be operated below (Fig. 4(a)) or above (Fig. 4(b)) the PT-symmetric threshold respectively. In addition, when the coupling detuning used in both waveguides is −2.325Γ, the system even becomes a non-symmetric one with much larger ratio (Fig. 4(c)).
Secondly, we present the field modes in the waveguides in Fig. 4(d–f) corresponding to Fig. 4(a–c), respectively. When the condition of the system is below the threshold (Fig. 4(a)), the PT symmetry condition is satisfied, the eigenvalues should be real, and the field modes should be symmetric. However, because of the imperfect PT symmetry condition, the eigenvalues have a very small imaginary part, and the two field modes are slightly asymmetric, as shown in Fig. 4(d). For the case of above the threshold (Fig. 4(b)), that is, the PT symmetry is broken, the eigenvalues become complex, where the imaginary part of the refractive index represents the gain or loss for each filed mode. Figure 4(e) illustrates that, the light modes become strongly asymmetric. For the non-symmetric one (Fig. 4(c)), the light modes are perfectly symmetric, which is shown in Fig. 4(f).
Thirdly, we show in Fig. 4(g–i) the propagation characteristics of the probe beam according to the condition of Fig. 4(a–c) respectively. When the condition of the system is below threshold, the probe beam oscillates periodically between the two waveguides, as show in Fig. 4(g). When the case is above the threshold, it can be seen from Fig. 4(h) that an exponentially growing mode occurs, which signifies the onset of PT symmetry breaking. For the passive waveguides, the probe beam also oscillates periodically (Fig. 4(i)), which is similar to the case of below threshold. However, the oscillation period in Fig. 4(i) is shorter than that of in Fig. 4(g).
In a practical system, it is very difficult to realize perfect PT-symmetric condition. So, we evaluate the degrees of asymmetry of the system by analyzing the asymmetry function of Δ(x) = n(x) − n*(−x), and find that the degree of asymmetry is very small. Therefore, it is manageable and has little impact on practical experiments.
PT-symmetric optical lattices—Type I
In this part, we first consider an 1D optical lattices of period Λx along x direction, and in each lattices the electron density of QWs is spatially modulated. The electron density in the ith trap can exhibit a Gaussian distribution \({N}_{i}(x,y)=N{e}^{-{(x-{x}_{i})}^{2}/{\sigma }_{x}^{2}}\), where N is the electron density of a homogeneous QW sample, xi is the ith trap center, and σx is the half width. The modulation of electron density of QWs can be achieved by using high order surface grating50,51. We further consider a pump laser field with periodically modulated intensity \({{\rm{\Omega }}}_{d}={{\rm{\Omega }}}_{d0}+\delta {{\rm{\Omega }}}_{d}\,\sin \,[2\pi (x-{x}_{i})/{{\rm{\Lambda }}}_{x}]\), which can be easily achieved in experiment by using an imperfect standing-wave (SW) field with unequal forward and backward components. We show the schematic diagram of QW system in Fig. 5. The probe field and coupling filed propagate along z direction, and the SW pump fields propagate along x direction.
First, we calculate the real χR and imaginary χI parts of the susceptibility as a function of Δp for varying value of Ωd by solving Eq. (3) without modulation of pump intensity or the electron density, and the corresponding results are shown in Fig. 6. It can be seen from red dotted line in Fig. 6(b) that when Ωd = 0.8Γ, a typical EIT with positive value (loss) is realized. With increasing value of Ωd, at both sides of the EIT windows, χI becomes positive (gain), which is knew as coherent Raman gain without population inversion [blue dashed line and black solid line in Fig. 6(b)]. Meanwhile, χR changes form positive dispersion to negative dispersion in the vicinity of EIT window, as shown in Fig. 6(a). The figures indicate that it is possible to realize PT symmetry by choosing suitable modulations.
(a) Real part χR and (b) imaginary part χI of the susceptibility as a function of probe detuning Δp for different pump Rabi frequency, Ωd = 0.8Γ (red dotted line), Ωd = Γ (blue dashed line), and Ωd = 1.2Γ (black solid line). The parameters are Ωc = Γ and Δc = Δd = 0. The other parameters are the same as in Fig. 3.
Then, with modulation of the pump intensity and electron density, we calculate the real (nR) and imaginary parts (nI) of the refractive index as a function of x and show the results in Fig. 7. Figure 7(a) shows the condition of Δp = 2.592 meV and δΩd = 0.1Γ. The value of Δp is chosen according to blue dashed line in Fig. 6(b), where χI has zero value around Δp = 2.592 meV, and thus it exhibits gain or absorption if Δp is slightly tuned away from this point. It can be seen that nR is an even function of the lattice position x, while nI is an odd function of lattice position x with balance gain and loss. Such results clearly indicate that PT symmetry is built in QW system by modulating the pump field and the electron density.
Real part nR and imaginary part nI of the refractive index as a function of position x for unchanged absolute value and different signs of Δp and δΩd, respectively, (a) Δp = 2.592 meV and δΩd = 0.1Γ, (b) Δp = 2.592 meV and δΩd = −0.1Γ, (c) Δp = −2.592 meV and δΩd = 0.1Γ, (d) Δp = −2.592 meV and δΩd = −0.1Γ. The parameters are Ωd0 = Γ. The other parameters are the same as in Fig. 6.
In Figs 7(b) and 6(d), we further show the cases for the same absolute value of Δp and δΩd as that of in Fig. 7(a), but with different sign of them. The results show that the PT-symmetric properties are maintained in all cases. However, the relation between nR and nI are different. It is found that positive (negative) value of Δp results in the positive (negative) value of nR and have no effect on the profile of nI. On the other hand, the sign of δΩd determines the position of gain and loss in each period, for instance, with positive (negative) value of δΩd, nI is gain (loss) in one half period −0.5 ≤ x/Λx ≤ 0. The understanding of impacts of the parameters on spatial features of the refractive index is important because of its potential application, such as asymmetric light diffraction52,53.
We next check what will happen if the absolute value of Δp and δΩd is changed. To do so, we plot 2D nR and nI as functions of x and Δp in Fig. 8(a,b), and as functions of x and δΩd in Fig. 8(c,d), respectively. The figures clearly show that nR and nI are modulated along x direction for varying Δp or δΩd. In addition, the profile of nI is much more sensitive to the parameters than that of nR. When Δp or δΩd is detuned from the value used in Fig. 7(a), through nR is maintained an even property, nI losses odd property, where the loss in one half period is not equal to the gain in the other half. Therefore, PT symmetry is destroyed in QW system because of the nonlinear response of nI to Δp or δΩd.
(a) Real part nR and (b) imaginary part nI of the refractive index as functions of position x and probe detuning Δp. The parameters are δΩd = 0.1Γ. (c) Real part nR and (d) imaginary part nI of the refractive index as functions of position x and modulation intensity δΩd. The parameters are Δp = 2.592 meV. The other parameters are the same as in Fig. 7.
Last, our goal is to realize 2D modulation of refractive index as functions of x and y, and this can be achieved by 2D modulation of the electron density and pump field, which are
We plot in Fig. 9(a,b) the top view of nR and nI as functions of x and y using the same parameters in Fig. 7(a), respectively. The results clearly show that nR is an even functions of x and y, while nI is an odd functions of x and y, therefore, 2D PT-symmetric optical lattices is realized in QW systems.
(a) Real part nR and (b) imaginary part nI of the refractive index as functions of position x and position y. The parameters are Δp = 2.592 meV and δΩd = 0.1Γ. The other parameters are the same as in Fig. 7.
PT-symmetric optical lattices—Type II
In this part, we show that PT symmetry can also be achieved by modulating both pump and coupling fields. The schematic diagram of the system is shown in Fig. 10, where the probe field propagates along z direction, and the SW coupling and SW pump fields propagate along x direction. Therefore, we have modulated coupling and pump fields, \({{\rm{\Omega }}}_{c}\) = \({{\rm{\Omega }}}_{c0}+\delta {{\rm{\Omega }}}_{c}\,\cos \,[2\pi (x-{x}_{i}){/{\rm{\Lambda }}}_{x}]\), \({{\rm{\Omega }}}_{d}\) = \({{\rm{\Omega }}}_{d0}+\delta {{\rm{\Omega }}}_{d}\,\sin \,[2\pi (x-{x}_{i}){/{\rm{\Lambda }}}_{x}]\).
First, we plot the real and imaginary parts of the refractive index (nR and nI) as a function of x in Fig. 11 for different cases. Figure 11(a) shows the condition of Δp = 2.592 meV and the modulation intensities δΩc = 0.3Γ and δΩd = 0.1Γ. It can be seen from figure that the PT symmetry is appears, where nR is an even function of x, and nI is an odd function of x. The relations of nR and nI can also be modified by changing the sign of the detuning Δp and the modulation intensities δΩc and δΩd, and the changing of their sign will not destroy the PT symmetry. There are eight kinds of combinations of these three parameters, corresponding to eight kinds of spatial refractive index, and we plot the other seven kinds in Fig. 11(b–h). It can be found that Δp, δΩc or δΩd have different impacts on the features of nR and nI. More specifically, Δp determines nR being positive or negative, and δΩc determines the monotonicity of nR, and δΩd determines the positions of the gain and loss in one period.
Real part nR and imaginary part nI of the refractive index as a function of position x for unchanged absolute value and different signs of Δp, δΩc and δΩd, respectively, (a) Δp = 2.525 meV, δΩc = 0.3Γ and δΩd = 0.1Γ, (b) Δp = 2.525 meV, δΩc = 0.3Γ and δΩd = −0.1Γ, (c) Δp = 2.525 meV, δΩc = −0.3Γ and δΩd = 0.1Γ, (d) Δp = 2.525 meV, δΩc = −0.3Γ and δΩd = −0.1Γ, (e) Δp = −2.525 meV, δΩc = 0.3Γ and δΩd = 0.1Γ, (f) Δp = −2.525 meV, δΩc = 0.3Γ and δΩd = −0.1Γ, (g) Δp = −2.525 meV, δΩc = −0.3Γ and δΩd = 0.1Γ, (h) Δp = −2.525 meV, δΩc = −0.3Γ and δΩd = −0.1Γ. The parameters are Ωc0 = 3Γ and Ωd0 = Γ. The other parameters are the same as in Fig. 6.
We also check the impact of absolute value of Δp, δΩc or δΩd on the properties of PT-symmetric system. Therefore, we plot 2D nR and nI as functions of x and Δp in Fig. 12(a,b), as functions of x and δΩd in Fig. 12(c,d), and as functions of x and δΩc in Fig. 12(e,f), respectively. The figures clearly show that nR and nI are modulated along x direction for varying Δp, δΩc or δΩd. However, when these parameters are detuned form the value used in Fig. 11(a), the system will loss the PT-symmetric properties.
(a) Real part nR and (b) imaginary part nI of the refractive index as functions of position x and probe detuning Δp. The parameters are δΩc = 0.3Γ and δΩd = 0.1Γ. (c) Real part nR and (d) imaginary part nI of the refractive index as functions of position x and modulation intensity δΩd. The parameters are Δp = 2.525 meV and δΩc = 0.3Γ. (e) Real part nR and (f) imaginary part nI of the refractive index as functions of position x and modulation intensity δΩc. The parameters are Δp = 2.525 meV and δΩd = 0.1Γ. The other parameters are the same as in Fig. 11.
Last, we show that it is possible to realize 2D PT-symmetric optical lattices by applying 2D modulation of the pump and coupling laser fields, which are
Using the same parameters in Fig. 11(a), we calculate nR and nI as functions of x and y, and show the tope view of nR and nI as functions of x and yin Fig. 13(a,b), respectively. The results indicate that 2D PT-symmetric optical lattices can be achieved, with nR being an even functions of x and y, and nI being an odd functions of x and y, respectively.
(a) Real part nR and (b) imaginary part nI of the refractive index as functions of position x and position y. The parameters are Δp = 2.525 meV, δΩc = 0.3Γ and δΩd = 0.1Γ. The other parameters are the same as in Fig. 11.
Conclusion
In conclusion, we have demonstrated that coherent asymmetric double semiconductor QWs can be an ideal candidate for studying PT symmetry. We have showed that PT-symmetric optical waveguides can be realized by using two coupling fields with different detuning, 1D and 2D PT-symmetric optical lattices can be realized by spatial modulation of pump laser fields and the electron density of QWs, or by spatial modulation of pump and coupling fields. In addition, the PT-symmetric properties realized in QW systems, such as the relation between the real and the imaginary parts of the complex refractive indices, can be controlled by changing the laser fields and the parameters of QWs.
References
Bender, C. M. & Boettcher, S. Real spectra in non-Hermitian Hamiltonians having PT symmetry. Physical Review Letters 80, 5243–5246 (1998).
El-Ganainy, R., Makris, K. G., Christodoulides, D. N. & Musslimani, Z. H. Theory of coupled optical PT-symmetric structures. Opt. Lett. 32, 2632–2634 (2007).
Ruter, C. E. et al. Observation of parity-time symmetry in optics. Nat Phys 6, 192–195 (2010).
Guo, A. et al. Observation of PT-Symmetry Breaking in Complex Optical Potentials. Physical Review Letters 103, 093902 (2009).
Regensburger, A. et al. Parity-time synthetic photonic lattices. Nature 488, 167–171 (2012).
Wimmer, M. et al. Observation of optical solitons in PT-symmetric lattices. Nature Communications 6, 7782 (2015).
Peng, B. et al. Parity-time-symmetric whispering-gallery microcavities. Nat Phys 10, 394–398 (2014).
Chang, L. et al. Parity-time symmetry and variable optical isolation in active-passive-coupled microresonators. Nat Photon 8, 524–529 (2014).
Feng, L. et al. Experimental demonstration of a unidirectional reflectionless parity-time metamaterial at optical frequencies. Nat Mater 12, 108–113 (2013).
Lin, Z. et al. Unidirectional Invisibility Induced by PT-Symmetric Periodic Structures. Physical Review Letters 106, 213901 (2011).
Longhi, S. PT-symmetric laser absorber. Physical Review A 82, 031801 (2010).
Chong, Y. D., Ge, L., Cao, H. & Stone, A. D. Coherent Perfect Absorbers: Time-Reversed Lasers. Physical Review Letters 105, 053901 (2010).
Hodaei, H., Miri, M.-A., Heinrich, M., Christodoulides, D. N. & Khajavikhan, M. Parity-time-symmetric microring lasers. Science 346, 975–978 (2014).
Feng, L., Wong, Z. J., Ma, R.-M., Wang, Y. & Zhang, X. Single-mode laser by parity-time symmetry breaking. Science 346, 972–975 (2014).
Jing, H. et al. PT-Symmetric Phonon Laser. Physical Review Letters 113, 053604 (2014).
Fleury, R., Sounas, D. & Alù, A. An invisible acoustic sensor based on parity-time symmetry. Nature Communications 6, 5905 (2015).
Musslimani, Z. H., Makris, K. G., El-Ganainy, R. & Christodoulides, D. N. Optical Solitons in PT Periodic Potentials. Physical Review Letters 100, 030402 (2008).
Longhi, S. Bloch Oscillations in Complex Crystals with PT-Symmetry. Physical Review Letters 103, 123601 (2009).
Liang, G. Q. & Chong, Y. D. Optical Resonator Analog of a Two-Dimensional Topological Insulator. Physical Review Letters 110, 203904 (2013).
Fejer, M. M., Yoo, S. J. B., Byer, R. L., Harwit, A. & HarrisJr, J. S. Observation of extremely large quadratic susceptibility at 9.6–10.8 um in electric-field-biased AlGaAs quantum wells. Physical Review Letters 62, 1041–1044 (1989).
Sirtori, C., Capasso, F., Sivco, D. L. & Cho, A. Y. Giant, triply resonant, third-order nonlinear susceptibility in coupled quantum wells. Physical Review Letters 68, 1010–1013 (1992).
Zhang, L. Electric field effect on the linear and nonlinear intersubband refractive index changes in asymmetrical semiparabolic and symmetrical parabolic quantum wells. Superlattices and Microstructures 37, 261–272 (2005).
Kuhn, K. J., Iyengar, G. U. & Yee, S. Free carrier induced changes in the absorption and refractive index for intersubband optical transitions in AlxGa1−xAs/GaAs/AlxGa1−xAs quantum wells. Journal of Applied Physics 70, 5010–5017 (1991).
Frogley, M. D., Dynes, J. F., Beck, M., Faist, J. & Phillips, C. C. Gain without inversion in semiconductor nanostructures. Nature Materials 5, 175 (2006).
West, L. C. & Eglash, S. J. First observation of an extremely large‐dipole infrared transition within the conduction band of a GaAs quantum well. Applied Physics Letters 46, 1156–1158 (1985).
Silvestri, L., Bassani, F., Czajkowski, G. & Davoudi, B. Electromagnetically induced transparency in asymmetric double quantum wells. The European Physical Journal B - Condensed Matter and Complex Systems 27, 89–102 (2002).
Phillips, M. & Wang, H. Electromagnetically induced transparency due to intervalence band coherence in a GaAs quantum well. Opt. Lett. 28, 831–833 (2003).
Yang, W.-X. & Lee, R.-K. Controllable entanglement and polarization phase gate in coupled double quantum-well structures. Opt. Express 16, 17161–17170 (2008).
Zhu, C. & Huang, G. Slow-light solitons in coupled asymmetric quantum wells via interband transitions. Physical Review B 80, 235408 (2009).
Luo, X. Q., Wang, D. L., Zhang, Z. Q., Ding, J. W. & Liu, W. M. Nonlinear optical behavior of a four-level quantum well with coupled relaxation of optical and longitudinal phonons. Physical Review A 84, 033803 (2011).
Sheng, J., Miri, M.-A., Christodoulides, D. N. & Xiao, M. PT-symmetric optical potentials in a coherent atomic medium. Physical Review A 88, 041803 (2013).
Li, H.-j, Dou, J.-p & Huang, G. PT symmetry via electromagnetically induced transparency. Opt. Express 21, 32053–32062 (2013).
Hang, C., Huang, G. & Konotop, V. V. PT Symmetry with a System of Three-Level Atoms. Physical Review Letters 110, 083604 (2013).
Wang, X. & Wu, J.-H. Optical PT-symmetry and PT-antisymmetry in coherently driven atomic lattices. Opt. Express 24, 4289–4298 (2016).
Zhang, Z. et al. Observation of Parity-Time Symmetry in Optically Induced Atomic Lattices. Physical Review Letters 117, 123601 (2016).
Kang, H. & Zhu, Y. Observation of Large Kerr Nonlinearity at Low Light Intensities. Physical Review Letters 91, 093601 (2003).
Schmidt, H. & Imamoglu, A. Giant Kerr nonlinearities obtained by electromagnetically induced transparency. Opt. Lett. 21, 1936–1938 (1996).
Sheng, J., Yang, X., Wu, H. & Xiao, M. Modified self-Kerr-nonlinearity in a four-level N-type atomic system. Physical Review A 84, 053820 (2011).
Abdullaev, F. K., Kartashov, Y. V., Konotop, V. V. & Zezyulin, D. A. Solitons in PT-symmetric nonlinear lattices. Physical Review A 83, 041805 (2011).
Hang, C., Zezyulin, D. A., Konotop, V. V. & Huang, G. Tunable nonlinear parity-time-symmetric defect modes with an atomic cell. Opt. Lett. 38, 4033–4036 (2013).
Ramezani, H., Kottos, T., El-Ganainy, R. & Christodoulides, D. N. Unidirectional nonlinear PT-symmetric optical structures. Physical Review A 82, 043803 (2010).
Roskos, H. G. et al. Coherent submillimeter-wave emission from charge oscillations in a double-well potential. Physical Review Letters 68, 2216–2219 (1992).
Nikonov, D. E., Imamoğlu, A., Butov, L. V. & Schmidt, H. Collective Intersubband Excitations in Quantum Wells: Coulomb Interaction versus Subband Dispersion. Physical Review Letters 79, 4633–4636 (1997).
Wu, J.-H. et al. Ultrafast All Optical Switching via Tunable Fano Interference. Physical Review Letters 95, 057401 (2005).
Scully, M. O. Enhancement of the index of refraction via quantum coherence. Physical Review Letters 67, 1855–1858 (1991).
Fleischhauer, M. et al. Resonantly enhanced refractive index without absorption via atomic coherence. Physical Review A 46, 1468–1487 (1992).
Zibrov, A. S. et al. Experimental Demonstration of Enhanced Index of Refraction via Quantum Coherence in Rb. Physical Review Letters 76, 3935–3938 (1996).
Yavuz, D. D. Refractive Index Enhancement in a Far-Off Resonant Atomic System. Physical Review Letters 95, 223601 (2005).
Proite, N. A., Unks, B. E., Green, J. T. & Yavuz, D. D. Refractive Index Enhancement with Vanishing Absorption in an Atomic Vapor. Physical Review Letters 101, 147401 (2008).
Dridi, K., Benhsaien, A., Zhang, J., Hinzer, K. & Hall, T. J. Narrow linewidth two-electrode 1560 nm laterally coupled distributed feedback lasers with third-order surface etched gratings. Opt. Express 22, 11 (2014).
Gao, F. et al. Study of gain-coupled distributed feedback laser based on high order surface gain-coupled gratings. Optics Communications 410, 936–940 (2018).
Liu, Y.-M., Gao, F., Fan, C.-H. & Wu, J.-H. Asymmetric light diffraction of an atomic grating with PT symmetry. Opt. Lett. 42, 4283–4286 (2017).
Shui, T., Yang, W.-X., Liu, S., Li, L. & Zhu, Z. Asymmetric diffraction by atomic gratings with optical PT symmetry in the Raman-Nath regime. Physical Review A 97, 033819 (2018).
Acknowledgements
This work was supported by National Natural Science Foundation of China (61774156 and 61761136009), External Cooperation Program of Chinese Academy of Sciences (181722KYSB20160005), Jilin Provincial Natural Science Foundation (20160520095JH and 20180519024JH), the program of China Scholarships Council (201604910385) and the Youth Innovation Promotion Association of Chinese Academy of Sciences (2018249).
Author information
Authors and Affiliations
Contributions
S.-C.T. proposed the research and carried out the calculations. R.-G.W. verified the results. All authors discussed the results and reviewed the manuscript.
Corresponding author
Ethics declarations
Competing Interests
The authors declare no competing interests.
Additional information
Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Tian, SC., Wan, RG., Wang, LJ. et al. Parity-time symmetry in coherent asymmetric double quantum wells. Sci Rep 9, 2607 (2019). https://doi.org/10.1038/s41598-019-39085-6
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41598-019-39085-6
This article is cited by
-
Azimuthal modulation of electromagnetically induced grating using structured light
Scientific Reports (2021)