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

Controlling quantum-dot light absorption and emission by a surface-plasmon field

Open Access Open Access

Abstract

The possibility for controlling both the probe-field optical gain and absorption, as well as photon conversion by a surface-plasmon-polariton near field is explored for a quantum dot located above a metal surface. In contrast to the linear response in the weak-coupling regime, the calculated spectra show an induced optical gain and a triply-split spontaneous emission peak resulting from the interference between the surface-plasmon field and the probe or self-emitted light field in such a strongly-coupled nonlinear system. Our result on the control of the mediated photon-photon interaction, very similar to the ‘gate’ control in an optical transistor, may be experimentally observable and applied to ultra-fast intrachip/interchip optical interconnects, improvement in the performance of fiber-optic communication networks, and developments of optical digital computers and quantum communications.

© 2014 Optical Society of America

1. Introduction

It is well known that photons inherently do not interact with each other. In classical electrodynamics, Maxwell equations are linear and cannot describe any photon-photon interaction. However, effective photon-photon coupling could exist in a mediated way, e.g., through their interactions with matter. Very recently, an experiment [1], which involves firing pairs of photons through an ultra-cold atomic gas, provided the evidence for an attractive interaction between photons to form so-called ‘molecules’ of light. In general, if the interaction between photons and matter is strong, the optical response of matter will become nonlinear and the resulting bandedge optical nonlinearities [2] will enable an effective photon-photon interaction [3]. An optical transistor [4] could be based on this idea, where ‘gate’ photons control the intensity of a light beam ‘source’. Optical transistors could be applied to speed up and improve the performance of fiber-optic communication networks. Here, all-optical digital signal processing and routing is fulfilled by arranging optical transistors in photonic integrated circuits, and the signal loss during the propagation could be compensated by inserting new types of optical amplifiers. Moreover, optical transistors are expected to play an important role in the development of an optical digital computer or quantum-encrypted communication.

Most previous research on optical properties of materials, including optical absorption, inelastic light scattering, and spontaneous emission, used a weak probe field as a perturbation to the studied system [5]. In this weak-coupling regime, the optical response of electrons depends only on the material characteristics [6], and, therefore, no photon-photon interaction is expected. However, the strong-coupling regime could be reached with help from microcavities, and the experimental effort on searching for polariton condensation (resulting from strong light-electron interaction) in semiconductors continues to produce results [79]. A general review of exciton-polariton condensation may be found in [10]. Successful demonstration of room-temperature polariton lasing without population inversion in semiconductor microcavities using both optical pumping [11, 12] and electrical injection [13, 14] have made it possible to achieve ultra-low lasing thresholds and very small emitter sizes comparable to the emitted wavelength. Semiconductor exciton-polariton nanolasers could advance intrachip and interchip optical interconnects by integrating them into semiconductor-based photonic chips. They might also have applications in medical devices and treatments, such as spatially selective illumination of individual neuron cells to locally control neuron firing activities in optogenetics and neuroscience, and near-field high-resolution imaging beyond the optical diffraction limit as well.

Theoretically, a big hurdle also exists for studying photon-photon interactions in the strong-coupling regime mainly due to intractable numerical calculations for systems with very strong nonlinearity. The existence of strong nonlinearity in such a system means that any perturbative theories, e.g. using bare electron states or linear response theory [5], become inadequate for describing both field and electron dynamics in this system. The presence of an induced polarization, regarded as a source term in Maxwell equations [15,16], from photo-excited electrons makes it impossible for us to solve the field equations by simply using finite-element analysis [17] or finite-difference-time-domain methods [18]. Although the semiconductor-Bloch equations [19] and density-matrix equations [5, 20], derived from many-body theory, are able to accurately capture the nonlinear optical response of electrons, the inclusion of pair scattering effects on both energy relaxation and optical dephasing precludes an analytic approach for seeking a solution of these equations. Consequently, there exist only very few theoretical studies on resonantly-coupled excitons to microcavity field [21], which depend heavily on computer simulations, and focus on simplified one-dimensional strongly-coupled microcavity systems.

Physically, not only high-quality microcavities [22] but also intense surface-plasmon near fields [23, 24] should be employed for reaching the strong-coupling goal in semiconductors. In this paper, we solve the self-consistent equations for strongly-coupled electromagnetic-field dynamics and electron quantum kinetics in a quantum dot above the surface of a thick metallic film. This is a situation that has not been fully explored so far from either a theoretical or experimental point of view. This is done based on finding an analytical solution to Green’s function [25,26] for a quantum dot coupled to a semi-infinite metallic material system, which makes it easy to calculate the effect of the induced polarization field as a source term in Maxwell equations. In our formalism, the strong light-electron interaction is reflected in the photon-dressed electronic states with a Rabi gap and in the feedback from the induced optical polarization of dressed electrons to the incident light. The formalism derived in this paper goes beyond the weak-coupling limit and deals with a much more realistic structure in the strong-coupling limit for the development of a surface-plasmon polariton laser with a very low threshold pumping. Our results clearly demonstrate the ability to control probe-field optical gain and absorption, and photon conversion by a surface-plasmon field with temperature-driven frequency detuning in such a nonlinear system led by dressed electron states, very similar to the ‘gate’ control in an optical transistor. These conclusions should be experimentally observable [27,28]. On the other hand, our numerical results also provide an example that demonstrates the so-called quantum plasmonics [29], where the properties of surface-plasmon polaritons and quantum-confined electrons are hybridized through near-field coupling.

In Sec. 2, we introduce our physics model and derive self-consistent equations for determining the coupled scattering dynamics of a surface-plasmon field and the quantum kinetics of electrons in quantum dots. Section 3 is devoted to a full discussion of our numerical results, including scattering and optical absorption of the surface-plasmon-polariton field by quantum dots, spontaneous emission, and the nonlinear optical response of the dressed electron states. Some concluding remarks are given in Sec. 4.

2. Model and theory

Our model system, as shown in Fig. 1, consists of a semi-infinite metallic material with a semiconductor quantum dot above its surface. A surface-plasmon-polariton (SPP) field is locally excited through a surface grating by normally-incident light. This propagating SPP field further excites an interband electron-hole (e-h) plasma in the quantum dot. The induced optical-polarization field of the photo-excited e-h plasma is strongly coupled to the SPP field to produce split degenerate e-h plasma and SPP modes with an anticrossing gap. A brief description of our self-consistent formalism was reported earlier [16]. In order to let readers follow easily the details of our model and formalism, we present here the full derivation of the Maxwell-Bloch numerical approach for an SPP field coupled to a photo-excited e-h plasma in the quantum dot.

 figure: Fig. 1

Fig. 1 Schematic illustration for a semi-infinite metal and a quantum dot above its surface at x3 = 0. Here, the surface-plasmon polariton (SPP) is locally excited by incident light with the help of a surface grating. The propagating SPP field further excites e-h pairs (plasmas) in the adjacent quantum dot. As a result, the optical-polarization field of the photo-excited e-h plasma is strongly coupled to the propagating SPP field to form split plasma-SPP modes with an anticrossing gap. Also, a probe-field is used for studying the photon dressing effect.

Download Full Size | PDF

2.1. General formalism

The Maxwell equation for a semi-infinite non-magnetic medium in position-frequency space can be written as [25]

××E(r;ω)εb(x3;ω)ω2c2E(r;ω)=ω2ε0c2𝒫loc(r;ω),
where E(r; ω) is the electric component of an electromagnetic field, H(r;ω)=(iωμ0)×E(r;ω) is the magnetic component of the electromagnetic field, r = (x1, x2, x3) is a three-dimensional position vector, ω is the angular frequency of the incident light, ε0, μ0 and c are the permittivity, permeability and speed of light in vacuum, 𝒫loc(r; ω) is an off-surface local polarization field generated by optical transitions of electrons in a quantum dot, and the position-dependent dielectric function is
εb(x3;ω)={εd,forx3>0εM(ω),forx3<0.
Here, εd characterizes for the semi-infinite dielectric material in the region x3 > 0, while εM(ω) is the dielectric function of the semi-infinite metallic material in the region x3 < 0. For the Maxwell equation in Eq. (1), we introduce the Green’s function 𝒢μν(r, r′; ω) that satisfies the following equation
μ[εb(x3;ω)ω2c2δλμ2xλxμ+δλμr2]𝒢μν(r,r;ω)=δλνδ(rr),
where r2=μ2xμ2 is the Laplace operator, δλμ represents the Kronecker delta, and the indices λ, μ = 1, 2, 3 indicate three spatial directions. Using the Green’s function defined in Eq. (3), we can convert the Maxwell equation in Eq. (1) into a three-dimensional integral equation
Eμ(r;ω)=Eμ(0)(r;ω)ω2ε0c2νd3r𝒢μν(r,r;ω)𝒫νloc(r;ω),
where Eμ(0)(r;ω) is a solution of the corresponding homogeneous equation
ν[εb(x3;ω)ω2c2δμν2xμxν+δμνr2]Eν(0)(r;ω)=0,
and the source term 𝒫νloc(r;ω) generally depends on the electric field in a nonlinear way and can be determined by the Bloch equation [2, 15].

2.2. Solving for Green’s function

For a semi-infinite medium, the Green’s function can be formally expressed by its Fourier transform

𝒢μν(r,r;ω)=d2k(2π)2eik(rr)gμν(k,ω|x3,x3),
where we have introduced the two-dimensional vectors r = (x1, x2) and k = (k1, k2). Substituting Eq. (6) into Eq. (3), we obtain
[εbω2c2k22+d2dx32k1k2ik1ddx3k1k2εbω2c2k12+d2dx32ik2ddx3ik1ddx3ik2ddx3εbω2c2k2][g11g12g13g21g22g23g31g32g33]=δ(x3x3)[100010001].
After a rotational transformation [25] is performed in k-space, i.e.,
fμν(k,ω|x3,x3)=μ,ν𝒮μμ(k)𝒮νν(k)gμν(k,ω|x3,x3),
where the rotational matrix 𝒮(k) is
𝒮(k)=1k[k1k20k2k1000k],
we acquire an equivalent version of Eq. (7)
[εbω2c2+d2dx320ikddx30εbω2c2k2+d2dx320ikddx30εbω2c2k2][f11f12f13f21f22f23f31f32f33]=δ(x3x3)[100010001].
To obtain the solution of Eq. (10), we need to employ both the finite-value boundary condition at x′3 = ±∞ and the continuity boundary condition at the x3 = 0 interface. This leads to the following five non-zero fμν (k, ω|x3, x′3) elements [25, 26]:
f22(k,ω|x3,x3)={(i2p)2ppd+peipdx3ipx3x3>0,x3<0(i2p)[eip|x3x3|pdppd+peip(x3+x3)],x3<0,x3<0(i2pd)[eipd|x3x3|+pdppd+peipd(x3+x3)],x3>0,x3>0(i2pd)2pdpd+peip(x3x3),x3<0,x3>0,
f13(k,ω|x3,x3)={ikc22εM(ω)ω2[2εM(ω)pdεM(ω)pd+εdp]eipdx3ipx3,x3>0,x3<0ikc22εM(ω)ω2[eip|x3x3|sgn(x3x3)+εM(ω)pdεdpεM(ω)pd+εdpeip(x3+x3)],x3<0,x3<0ikc22εdω2[eipd|x3x3|sgn(x3x3)+εM(pd)pdεdpεM(ω)pd+εdpeipd(x3+x3)],x3>0,x3>0ikc22εdω2[2εdpεM(ω)pd+εdp]eipx3+ipdx3,x3<0,x3>0
f33(k,ω|x3,x3)={ik2c2ω2[1εM(ω)pd+εdp]eipdx3ipx3,x3>0,x3<0c2εM(ω)ω2δ(x3x3)ik2c22pεM(ω)ω2[eip|x3x3|εM(ω)pdεdpεM(ω)pd+εdpeip(x3+x3)],x3<0,x3<0c2εdω2δ(x3x3)ik2c22pdεdω2[eipd|x3x3|+εM(ω)pdεdpεM(ω)pd+εdpeipd(x3+x3)],x3>0,x3>0ik2c2ω2[1εM(ω)pd+εdp]eipx3+ipdx3,x3<0,x3>0
f11(k,ω|x3,x3)={ipdpc2ω2[1εM(ω)pd+εdp]eipdx3ipx3,x3>0,x3<0ipc22εM(ω)ω2[eip|x3x3|+εM(ω)pdεdpεM(ω)pd+εdpeip(x3+x3)],x3<0,x3<0ipdc22εdω2[eipd|x3x3|εM(ω)pdεdpεM(ω)pd+εdpeipd(x3+x3)],x3>0,x3>0ipdc22εdω2[2εdpεM(ω)pd+εdp]eipx3+ipdx3,x3<0,x3>0
f31(k,ω|x3,x3)={ikc2ω2[pεM(ω)pd+εdp]eipdx3ipx3,x3>0,x3<0ikc22εM(ω)ω2[eip|x3x3|sgn(x3x3)εM(ω)pdεdpεM(ω)pd+εdpeip(x3+x3)],x3<0,x3<0ikc22εdω2[eipd|x3x3|sgn(x3x3)εM(ω)pdεdpεM(ω)pd+εdpeipd(x3+x3)],x3>0,x3>0ikc2ω2[pdεM(ω)pd+εdp]eipx3+ipdx3,x3<0,x3>0
where sgn(x) is the sign function,
pd(k,ω)=εdω2c2k2,
p(k,ω)=εM(ω)ω2c2k2,
Im[pd(k, ω)] ≥ 0 and Im[p(k, ω)] ≥ 0. In addition, from these non-zero fμν (k, ω|x3, x′3) functions, we obtain
gμν(k,ω|x3,x3)=μ,νfμν(k,ω|x3,x3)𝒮μμ(k)𝒮νν(k),
which can be substituted into Eq. (6) to calculate the Green’s function 𝒢μν(r, r′; ω) in position space.

2.3. Local polarization field

In order to find the explicit field dependence in 𝒫loc(r; ω), we now turn to the study of electron dynamics in a quantum dot. Here, the optical-polarization field 𝒫loc(r; ω) plays a unique role on bridging the classical Maxwell equations for electromagnetic fields to the quantum-mechanical Schrödinger equation for electrons. The electron dynamics in photo-excited quantum dots can be described quantitatively by the so-called semiconductor Bloch equations [3032]. These generalize the well-known optical Bloch equations in two aspects including the incorporation of electron scattering with impurities, phonons and other electrons as well as many-body effects on dephasing in the photo-induced optical coherence.

The physical system considered in this paper is illustrated in Fig. 2, where we assume two levels for electrons and holes, respectively, in a quantum dot. These two energy levels of both electrons and holes are efficiently coupled by phonon scattering at high temperatures. Additionally, the lowest electron and hole energy levels are optically coupled to each other by an incident SPP field to form the dressed states of excitons. The SPP-controlled optical properties of quantum-dot excitons can either probed by a plane-wave field or seen from the spontaneous emission of excitons.

 figure: Fig. 2

Fig. 2 Schematic illustration for a system incorporating the generation of quantum-dot excitons by a SPP field with frequency ωsp and probed by a plane-wave field Ep with frequency ωp. Here, G is the energy bandgap of the host semiconductor, εe and εjh stand for the energy levels of electrons and holes, respectively, with , j = 1, 2, ···. In addition, ωem represents the frequency of spontaneous emission from photo-excited excitons, and the ground states of electrons and holes are coupled to their first excited states by lattice phonons at finite temperatures.

Download Full Size | PDF

For photo-excited spin-degenerated electrons in the conduction band, the semiconductor Bloch equations with = 1, 2, ··· are given by

dnedt=2h¯jIm[(Yj)*(,jehYjV,j;j,eh)]+net|relδ,1spn1en1h,
where sp is the spontaneous emission rate and ne represents the electron level population. In Eq. (19), the term marked ‘rel’ is the non-radiative energy relaxation for ne, and the Yj, ,jeh, and V,j;j,eh terms are given later in the text.

Similarly, for spin-degenerate holes in the valence band, the semiconductor Bloch equations with j = 1, 2, ··· are found to be

dnjhdt=2h¯Im[(Yj)*(,jehYjV,j;j,eh)]+njht|relδj,1spn1en1h,
where njh stands for the hole energy level population. Again, the non-radiative energy relaxation for njh is incorporated in Eq. (20). Moreover, we know from Eqs. (19) and (20) that
Ne(t)=2ne(t)=2jnjh(t)=Nh(t),
where Ne(t) and Nh(t) are the total number of photo-excited electrons and holes, respectively, in the quantum dot at time t.

Finally, for spin-averaged e-h plasmas, the induced interband optical coherence, which is introduced in Eqs. (19) and (20), with j = 1, 2, ··· and = 1, 2, ··· satisfies the following equations,

ih¯ddtYj=[ε¯e(ω)+ε¯jh(ω)h¯(ω+iγ0)]Yj+(1nenjh)(,jehYjV,j;j,eh)+Yj[j1nj1h(Vj,j1;j1,jhhVj,j1;j,j1hh)1n1eV1,j;j,1eh]+Yj[1n1e(V,1;1,eeV,1;,1ee)j1nj1hV,j1;j1,eh],
where h̄γ0 = h̄γeh + h̄γext is the total energy-level broadening due to both the finite carrier life-time and the loss of an external evanescent field, ω is the frequency of the external field, and ε¯e(ω) and ε¯jh(ω) are the kinetic energies of dressed single electrons and holes, respectively (see Appendix A with α = 1). In Eq. (22), the diagonal dephasing (γ0) of Yj, the renormalization of interband Rabi coupling ( YjV,j;j,eh), the renormalization of electron and hole energies (third and fourth terms on the right-hand side), as well as the exciton binding energy, are all taken into consideration. Since the e-h plasmas are independent of spin index in this case, they can be excited by both left-circularly and right-circularly polarized light. The off-diagonal dephasing of Yj has been neglected due to low carrier density.

The steady-state solution to Eq. (22), i.e. under the condition of dYj/dt=0, is found to be

Yj(t|ω)=[1ne(t)njh(t)h¯(ω+iγ0)h¯Ω¯,jeh(ω|t)],jeh(t),
where the photon and Coulomb renormalized interband energy-level separation h¯Ω¯,jeh(ω|t) is given by
h¯Ω¯,jeh(ω|t)=ε¯e(ω|t)+ε¯jh(ω|t)V,j;j,eh+1n1e(t)(V,1;1,eeV,1;,1ee)+j1nj1h(t)(Vj,j1;j1,jhhVj,j1;j,j1hh)1n1e(t)V1,j;j,1ehj1jnj1h(t)V,j1;j1,eh.
The steady-state solution in Eq. (23) can be substituted into Eqs. (19) and (20) above.

The Coulomb interaction matrix elements introduced in Eqs.(19), (20) and (22) are defined as

V1,2;3,4ee=d3rd3r[ψ1e(r)]*[ψ2e(r)]*e24πε0εb|rr|ψ3e(r)ψ4e(r)=e28π2ε0εbd2q1,4e(q)2,3e(q)(1q+qs)=(V1,2;3,4ee)*,
Vj1,j2;j3,j4hh=d3rd3r[ψj1h(r)]*[ψj2h(r)]*e24πε0εb|rr|ψj3h(r)ψj4h(r)=e28π2ε0εbd2qj1,j4h(q)j2,j3h(q)(1q+qs)=(Vj1,j2;j3,j4hh)*,
V,j;j,eh=d3rd3r[ψe(r)]*[ψjh(r)]*e24πε0εb|rr|ψjh(r)ψe(r)=e28π2ε0εbd2q,e(q)j,jh(q)(1q+qs)=(V,j;j,eh)*,
where the static screening length 1/qs at temperatures (kBTEF) is determined from
qs(t)=e24ε0εb𝒮kBT[Ne(t)+Nh(t)].
Here, 𝒮 is the cross-sectional area of a quantum dot, T is the lattice temperature, ψe(r) and ψjh(r) are the envelope wave-functions of electrons and holes in a quantum dot (see Appendix A), and εb is the average dielectric constant of the host semiconductor. The two dimension-less form factors (see Appendix A) introduced in Eqs. (25)(27) for electrons and holes due to quantum confinement by a quantum dot are defined by
,e(q)=eq0d2r[ψe(r)]*eiqrψe(r)=[,e(q)]*,
j,jh(q)=eq0d2r[ψjh(r)]*eiqrψjh(r)=[j,jh(q)]*,
where 0 is the thickness of a disk-like quantum dot. In addition, the matrix elements employed in Eqs. (19), (20) and (22) for the Rabi coupling between photo-excited carriers and an evanescent external field E(r; t) = θ(t)E(r; ω)eiωt are given by
,jeh(t)=δ,1δj,1θ(t)[E,jeh(ω)dc,v],
where θ(x) is a unit step function, the static interband dipole moment dc,v (see Appendix A) is
dc,v=d3r[uc(r)]*ruv(r)=dc,v*,
uc(r) and uv(r) are the Bloch functions associated with conduction and valence bands at the Γ-point in the first Brillouin zone of the host semiconductor, and the effective electric field coupled to the quantum dot is
E,jeh(ω)=d3r[ψe(r)]*E(r;ω)[ψjh(r)]*.

The Boltzmann-type scattering term [33] for non-radiative electron energy relaxation in Eq. (19) is

net|rel=𝒲(in)(1ne)𝒲(out)ne,
where the microscopic scattering-in and scattering-out rates are calculated as
𝒲(in)=2πh¯|V,ep|2ne{𝒩ph(Ω0)[h¯Γph/π(ε¯eε¯eh¯Ω0)2+h¯2Γph2]+[𝒩ph(Ω0)+1][h¯Γph/π(ε¯eε¯e+h¯Ω0)2+h¯2Γph2]}+2πh¯j,j|V,j;j,eh|2(1njh)njhne[h¯γeh/π(ε¯e+ε¯jhε¯eε¯jh)2+h¯2γeh2],
𝒲(out)=2πh¯|V,ep|2(1ne){𝒩ph(Ω0)[h¯Γph/π(ε¯eε¯eh¯Ω0)2+h¯2Γph2]+[𝒩ph(Ω0)+1][h¯Γph/π(ε¯eε¯e+h¯Ω0)2+h¯2Γph2]}+2πh¯j,j|V,j;j,eh|2(1ne)(1njh)njh[h¯γeh/π(ε¯e+ε¯jhε¯eε¯jh)2+h¯2γeh2].
Here, the primed summations in Eqs. (35) and (36) exclude the terms satisfying either j = j′ or ℓ′ = , 𝒩ph0) = [exp(Ω0/kBT) − 1]−1 is the Bose function for the thermal-equilibrium phonons, and Ω0 and Γph are the frequency and lifetime of longitudinal-optical phonons in the host semiconductor. Similarly, the Boltzmann-type scattering term for hole non-radiative energy relaxation in Eq. (20) is
njht|rel=𝒲¯j(in)(1njh)𝒲¯j(out)njh,
where the scattering-in and scattering-out rates are
𝒲¯j(in)=2πh¯j|Vj,jhp|2njh{𝒩ph(Ω0)[h¯Γph/π(ε¯jhε¯jhh¯Ω0)2+h¯2Γph2]+[𝒩ph(Ω0)+1][h¯Γph/π(ε¯jhε¯jh+h¯Ω0)2+h¯2Γph2]}+2πh¯,j|V,j;j,eh|2(1ne)njhne[h¯γeh/π(ε¯e+ε¯jhε¯eε¯jh)2+h¯2γeh2],
𝒲¯j(out)=2πh¯j|Vj,jhp|2(1njh){𝒩ph(Ω0)[h¯Γph/π(ε¯jhε¯jhh¯Ω0)2+h¯2Γph2]+[𝒩ph(Ω0)+1][h¯Γph/π(ε¯jhε¯jh+h¯Ω0)2+h¯2Γph2]}+2πh¯,j|V,j;j,eh|2(1ne)(1njh)ne[h¯γeh/π(ε¯e+ε¯jhε¯eε¯jh)2+h¯2γeh2],
and again the primed summations in Eqs. (38) and (39) exclude the terms satisfying either j′ = j or = ℓ′. The coupling between the longitudinal-optical phonons and electrons or holes in Eqs. (35), (36), (38) and (39) are
|V,ep|2=e2h¯Ω08π2ε0(1ε1εs)d2q|,e(q)|2(1q+qs),
|V,hp|2=e2h¯Ω08π2ε0(1ε1εs)d2q|j,jh(q)|2(1q+qs),
where ε and εs are the high-frequency and static dielectric constants of the host polar semi-conductor.

By generalizing the Kubo-Martin-Schwinger relation [20], the time-dependent spontaneous emission rate, sp(t), introduced in Eqs. (19) and (20), can be expressed as

sp(t)=|dc,v(t)|2ε0εb|d3rψ1e(r)ψ1h(r)|20dωθ[h¯ωc(t)ε¯1e(ω|t)ε¯1h(ω|t)]×h¯ωρ0(ω){h¯γeh[h¯ωc(t)ε¯1e(ω|t)ε¯1h(ω|t)]2+h¯2γeh2},
where
|dc,v(t)|2=e2h¯22m0G(T)[1+c(t)G(T)](m0me*1),
G(T) = G(0) − 5.41 × 10−4 T2/(T + 204) (in units of eV) is the energy bandgap of the host semiconductor, ρ0(ω) = ω2/c3π2 is the density-of-states of spontaneously-emitted photons in vacuum, m0 is the free electron mass, me* is the effective mass of electrons, and the Coulomb renormalization of the energy bandgap c(t) is found to be
c(t)=1n1e(t)(V1,1;1,1eeV1,1;1,1ee)+j1nj1h(t)(V1,j1;j1,1hhV1,j1;1,j1hh)1n1e(t)V1,1;1,1ehj1nj1h(t)V1,j1;j1,1eh[1n1e(t)n1h(t)]V1,1;1,1eh.
In Eq. (44), the first two terms are associated with the Hartree-Fock energies of electrons and holes, while the remaining terms are related to the exciton binding energy.

Finally, the photo-induced interband optical polarization 𝒫loc(r; ω), which is related to the induced interband optical coherence, by dressed electrons in the quantum dot is given by [2]

𝒫loc(r;ω)=2|ξ(r)|2dc,v{d3rψ1e(r)ψ1h(r)}×1h¯limt{1n1e(t)n1h(t)ω+iγ0Ω¯1,1eh(ω|t)}1,1eh(t),
where dc,v = dc,v êd represents the interband dipole moment [see Eq. (32)], êd is the unit vector of the dipole moment, and |ξ(r)|2 comes from the confinement of a quantum dot.

2.4. Self-consistent field equation

Since the wavelength of the incident light is much larger than the size of a quantum dot, we can treat the quantum dot, which is excited resonantly by the incident light, as a point dipole at r = r0 = (0, 0, z0), i.e. we can assume 𝒫loc(r′; ω) = 𝒫loc(ω)δ(r′r0) in Eq. (4) to neglect its geometry effect. This greatly simplifies the calculation and gives rise to

Eμ(r;ω)=Eμ(0)(r;ω)ω2ε0c2ν𝒢μν(r,r0;ω)𝒫νloc(ω),
where
𝒫loc(ω)=2dc,v{d3rψ1e(r)ψ1h(r)}×1h¯limt{1n1e(t)n1h(t)ω+iγ0Ω¯1,1eh(ω|t)}1,1eh(t),
1,1eh(t)=θ(t)[E(r0;ω)dc,v]{d3rψ1e(r)ψ1h(r)}*.
Substituting Eqs. (47) and (48) into Eq. (46), we get the following nonlinear equations for the electromagnetic field
Eμ(r;ω)=Eμ(0)(r;ω)+2ω2ε0c2[E(r0;ω)dc,v]dc,v|d3rψ1e(r)ψ1h(r)|2×1h¯limt{1n1e(t)n1h(t)ω+iγ0Ω¯1,1eh(ω|t)}ν𝒢μν(r,r0;ω)e^dν,
where the quantum-dot level populations ne(t) and njh(t) depend nonlinearly on E(r0; ω) in the strong-coupling regime.

If the electromagnetic field is not very strong, we can neglect the pumping effect. In this linear-response regime, we can write down the electron and hole populations in a thermal-equilibrium state [without solving Eqs. (19) and (20)]

ne(t)f0(ε¯e)1exp[(ε¯eμe)/kBT]+1,
njh(t)f0(ε¯jh)1exp[(ε¯jhμh)/kBT]+1,
where f0(x) is the Fermi function, and μe and μh are the chemical potentials of electrons and holes, respectively, determined by Eq. (21). As a result of Eqs. (50) and (51), we get from Eq. (49) the linearized self-consistent field equation at r = r0
ν𝒜μν(r0;ω)Eν(r0;ω)=Eμ(0)(r0;ω)
with
𝒜μν(r0;ω)=δμν2ω2ε0c2h¯[1f0(ε¯1e)f0(ε¯1h)ω+iγ0Ω¯1,1eh(ω)]|d3rψ1e(r)ψ1h(r)|2×dc,v2[e^dνν1𝒢μν1(r0,r0;ω)e^dν1],
where, according to Eq. (6), we have
𝒢μν(r0,r0;ω)=d2k(2π)2gμν(k,ω|z0,z0).
The solution E(r0; ω) of the linear-matrix equation in Eq. (52) can be substituted into Eq. (49) to yield the spatial distribution of the electromagnetic field E(r; ω) at all positions other than r = r0, i.e.,
Eμ(r;ω)=Eμ(0)(r;ω)+2ω2ε0c2h¯[ν,νe^dν𝒜νν1(r0;ω)Eν(0)(r0;ω)]×|d3rψ1e(r)ψ1h(r)|2dc,v2[1f0(ε¯1e)f0(ε¯1h)ω+iγ0Ω¯1,1eh(ω)]ν1𝒢μν1(r,r0;ω)e^dν1.

In order to find the coupled e-h plasma and plasmon dispersion relation ω = Ωex–pl(k), we Fourier transform both E(r; ω) and E(0)(r; ω) in Eq. (46) with respect to r. This leads to

Eμ(k,ω|x3)=Eμ(0)(k,ω|x3)ω2ε0c2νgμν(k,ω|x3,z0)𝒫νloc(ω).
After setting x3 = z0 in Eq. (56), we get
ν{δμν2ω2ε0c2h¯[1f0(ε¯1e)f0(ε¯1h)ω+iγ0Ω¯1,1eh(ω)]|d3rψ1e(r)ψ1h(r)|2dc,v2×[e^dνν1gμν1(k,ω|z0,z0)e^dν1]}Eν(k,ω|z0)=Eμ(0)(k,ω|z0).
Here, the vanishing of the determinant of the coefficient matrix in Eq. (57) determines the coupled e-h plasma and plasmon dispersion relation ω = Ωex–pl(k). We emphasize that the assumption of thermal-equilibrium states for electrons and holes is just for obtaining analytical expressions. Therefore, some qualitative conclusions can be drawn for guidance from these analytical solutions. Our numerical results, however, are based on the non-thermal-equilibrium steady states calculated after solving self-consistently the coupled Maxwell-Bloch equations.

By assuming an incident SPP field within the x1x2-plane, we can write

E(0)(r;ωsp)=Espeik0(ωsp)D0cωsp[ik^0β3(k0,ωsp)x^3k0(ωsp)]eik0(ωsp)xeβ3(k0,ωsp)x3,
where x = {x1, x2}, 0 and 3 are the unit vectors in the k0 = k0(ωsp){cosθ0, sinθ0} and x3 directions, Esp is the field amplitude, ωsp is the field frequency, θ0 is the angle of the incident SPP field with respect to the x1 direction, D0 = {−xg, −yg} is the position vector of the surface grating, and the two wave numbers are
k0(ωsp)=ωspcεdεM(ωsp)εd+εM(ωsp),
β3(k0,ωsp)=k02(ωsp)ωsp2c2,
with Re[k0(ωsp)] ≥ 0 and Re[β3(k0, ωsp)] ≥ 0. Here, the in-plane wave number k0 is produced by the surface-grating diffraction of the p-polarized normally-incident light, which in turn determines the resonant frequency ω of the SPP mode. Equation (59) stands for the full dispersion relation of the SPP field, including both radiative and non-radiative parts. From Eq. (58), it is easy to find its Fourier transformed expression
E(0)(k,ωsp|z0)=δ(kk0)Espeik0(ωsp)D0(2π)2cωsp×[ik^0β3(k0,ωsp)x^3k0(ωsp)]eβ3(k0,ωsp)z0.

2.5. Quantum-dot absorption

On the basis of the electromagnetic field E(r0; ω) at the quantum dot, we are able to compute the time-resolved nonlinear interband absorption coefficient of dressed electrons in a quantum dot for the SPP field [34]. In this case, we find

β0(ω;t)=ωεbnspp(ω;t)c[1exp(h¯ω/kBT)1+1]Im[αspp(ω;t)],
where αspp(ω; t) is the complex Lorentz function given by
Im[αspp(ω;t)]=θ(t)(2ε0εb𝒱|E(0)(r0;ω)|2)|E(r0;ω)dc,v|2|d3rψ1e(r)ψ1h(r)|2×[1n1e(t)n1h(t)]{[A2(ω;t)B2(t)]2+4h¯2γ02A2(ω;t)[A2(ω;t)+B2(t)]2+4h¯2γ02A2(ω;t)}[h¯γ0Δ2(ω;t)+h¯2γ02],
Re[αspp(ω;t)]=θ(t)(2ε0εb𝒱|E(0)(r0;ω)|2)|E(r0;ω)dc,v|2|d3rψ1e(r)ψ1h(r)|2×[1n1e(t)n1h(t)]{A4(ω;t)B4(t)]2[A2(ω;t)+B2(t)]2}[Δ(ω;t)Δ2(ω;t)+h¯2γ02],
and the scaled refractive index function nspp(ω; t) can be calculated by
nspp(ω;t)=12(1+Re[αspp(ω;t)]+{1+Re[αspp(ω;t)]}2+{Im[αspp(ω;t)]}2)1/2.
In Eqs. (63) and (64), the dressed-state effects on both the level population and dipole moment have been included. In addition, we have introduced the following notations in Eqs. (63) and (64)
Δ(ω;t)=[G(T)+ε1e+ε1hh¯ω]2+4|1,1eh(t)|2,
A2(ω;t)=[h¯ωG(T)ε1eε1h+Δ(ω;t)]2,B2(t)=4|1,1eh(t)|2.

2.6. Probing quantum-dot dressed states

We are also able to compute the time-resolved linear interband absorption coefficient of electrons, dressed by the SPP field, for a weak probe field (not the strong SPP field) on the basis of the electromagnetic field E(r0; ω) at the quantum dot calculated above [34]. Assuming a spatially-uniform probe field Ep(t) = θ(tτ) Ep ept where τ is the delay time, the probefield absorption coefficient βabs(ωp; t) of the lowest dressed state is given by Eq. (62) with the replacements of ω, nspp, and αspp by ωp, npf, and αpf, respectively, where

αpf(ωp;t)=θ(tτ)(2ε0εb𝒱|Ep|2h¯)|Epdc,v|2|d3rψ1e(r)ψ1h(r)|2[1n1e(t)n1h(t)]×{A2(ω;t)B2(t)[A2(ω;t)+B2(t)]2}{A2(ω;t)ωp+iγehΩ¯1,1eh(ω|t)B2(t)ωp+iγehΩ¯1,1eh(ω+|t)},
npf(ωp;t)=12(1+Re[αpf(ωp;t)]+{1+Re[αpf(ωp;t)]}2+{Im[αpf(ωp;t)]}2)1/2.
Here, using Eq. (24) we have
h¯Ω¯1,1eh(ω±|t)=h¯ω±(t)[1n1e(t)n1h(t)]V1,1;1,1eh+1n1e(t)(V1,1;1,1eeV1,1;1,1ee)+j1nj1h(t)(V1,j1;j1,1hhV1,j1;1,j1hh)1n1e(t)V1,1;1,1ehj1nj1h(t)V1,j1;j1,1eh,
and
h¯ω±(t)=h¯ω±Δ(ω;t).

Moreover, the time-resolved photoluminescence spectrum 𝒫em(ω′; t) of dressed electrons in the quantum dot is proportional to

𝒫em(ω;t)|dc,v|2ε0εb0n1e(t)n1h(t)h¯γeh{1[A2(ω;t)+B2(t)]2}|d3rψ1e(r)ψ1h(r)|2h¯ωρ0(ω)×{A2(ω;t)B2(t)[h¯ωc(t)h¯ω(t)]2+h¯2γeh2+A2(ω;t)B2(t)[h¯ωc(t)h¯ω+(t)]2+h¯2γeh2+A4(ω;t)+B4(t)[h¯ωc(t)h¯ω]2+h¯2γeh2}.

3. Numerical results and discussions

3.1. Results for the dynamics of an SPP field

In the first part of our numerical calculations, we have taken: 0 = 100 Å, Ly = 100 Å, me*=0.067m0, mh*=0.62m0, θ0 = 45°, xg = yg = 610 Å, εb = 12, εs = 11, ε = 13, Ω0 = 36 meV and h̄γeh = Γph = h̄γ0. The silver plasma frequency is 13.8 × 1015 Hz and the silver plasma dephasing is 0.1075×1015 Hz. The energy gap EG of the quantum-dot host material is 1.927 eV at T = 300 K. Other parameters, including T, Esp, Lx, h̄γ0, z0 and εd, will be directly indicated in the figures.

Figure 3 presents the quantum dot absorption coefficient β0(ωsp) for an SPP field, the scattered field |EtotEsp| of the SPP field, and the energy-level occupations for electrons nℓ,e and holes nj,h with , j = 1, 2 as functions of frequency detuning Δh̄ωsph̄ωsp − (EG +ε1,e +ε1,h). A dip is observed at resonance Δh̄ωsp = 0 in the upper-left panel, which appears to become deeper with decreasing amplitude Esp of the SPP field in the strong-coupling regime due to a decrease in the saturated absorption. However, this dip completely disappears when Esp drops to 25 kV/cm in the weak-coupling regime due to the suppression of the photon-dressing effect, which is accompanied by an order of magnitude increase in the absorption-peak strength. The dip in the upper-left panel corresponds to a peak in the scattered field, as can be seen from the upper-right panel of the figure. The scattered field increases with frequency detuning away from resonance, corresponding to the decreasing absorption. As a result, a minimum appears on each side of the resonance in the scattered field in the strong-coupling regime. The Maxwell-Bloch equations couple the field dynamics outside of a quantum dot with the electron dynamics inside the dot. At Esp = 125 kV/cm in the lower-right panel, we find peaks in the energy-level occupations at resonance, which are broadened by the finite carrier lifetime as well as the optical power of the SPP field. Moreover, jumps in the energy-level occupations can be seen at resonance due to Rabi splitting of the energy levels in the dressed electron states. The effect of resonant phonon absorption also plays a significant role in the finite value of n2,e with energy-level separations ε2,eε1,eΩ0. However, as Esp decreases to 25 kV/cm in the lower-left panel, peaks in the energy-level occupations are greatly sharpened and negatively shifted due to the suppression of the broadening from the optical power and the excitonic effect, respectively. Additionally, jumps in the energy-level occupations become invisible because the Rabi-split energy gap in this case is much smaller than the energy-level broadening from the finite lifetime of electrons (i.e. severely damped Rabi oscillations between the first electron and hole levels).

 figure: Fig. 3

Fig. 3 Optical absorption coefficients β0(ωsp) (upper-left panel) and scattering field |EtotEsp| at the quantum dot (upper-right panel), as well as the energy-level occupations for electrons nℓ,e and holes nj,h (lower panels) with , j = 1, 2, as functions of the frequency detuning Δh̄ωsph̄ωsp − (EG +ε1,e +ε1,h). Here, the results for various amplitudes Esp of an SPP field with frequency ωsp are presented in the upper panels, along with a comparison of the energy-level occupations for Esp = 25 and 125 kV/cm in the lower panels. The label ×0.1 in the upper-left panel indicates that the result is multiplied by a factor of 0.1.

Download Full Size | PDF

We know that a decrease in temperature T gives rise to an increase in the crystal bandgap EG. On the other hand, the localization of an SPP field (i.e. an exponential decay of the field strength on either side of a metallic surface) is greatly enhanced when the SPP frequency ωsp approaches that of a surface plasmon. As a result, the field at the quantum dot is expected to decrease as T is reduced. This gives rise to a higher absorption coefficient for a lower temperature, as shown in the upper-left panel of Fig. 4. Interestingly, although the suppressed absorption coefficient can be seen from β0(ωsp) for high SPP-field amplitudes, as shown by Eq. (63), from the upper-right panel of this figure we find the resonant peak at h̄ωsp = EG + ε1,e + ε1,h initially increases with T but then decreases with T at room temperature. This subtle difference demonstrates the effect of reduced phonon absorption at T = 77 K on the resonant scattered field by the factor 1 − ne(t) − nh(t) in Eq. (49). Moreover, the strong effect of the suppressed optical-phonon absorption between two electron energy levels at 77 K is clearly demonstrated in the lower panels of Fig. 4, where the level occupation n2,e becomes negligible at T = 77 K in comparison with that at T = 300 K.

 figure: Fig. 4

Fig. 4 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as nℓ,e and holes nj,h (lower panels), as functions of Δh̄ωsp. Here, the results for three different temperatures T = 300, 175 and 77 K are displayed in the upper panels, along with a comparison of nℓ,e and holes nj,h for T = 300 and 77 K in the lower panels.

Download Full Size | PDF

The electron thermal dynamics due to phonon absorption has been demonstrated in Fig. 4 for various temperatures. In Fig. 5, we present the electron dynamics resulting from the optical dephasing, due to the finite lifetime of electrons, at different energy-level broadenings h̄γ0. As h̄γ0 is increased from 3 meV to 7 meV, the dip in β0(ωsp) at resonance is suppressed, leading to a single peak with a reduced strength and an increased width, as shown in the upper-left panel of the figure. This increase in the resonant absorption is further accompanied by an enhanced resonant peak for the scattered field in the upper-right panel of this figure. As expected, the energy-level occupations at h̄γ0 = 7 meV become much broader than those at h̄γ0 = 3 meV, as displayed in the lower two panels of the figure.

 figure: Fig. 5

Fig. 5 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as nℓ,e and holes nj,h (lower panels), as functions of Δh̄ωsp. Here, the results with different energy-level broadening h̄γ0 = 3, 5 and 7 meV are shown in the upper panels, along with a comparison of nℓ,e and holes nj,h for h̄γ0 = 3 and 7 meV in the lower panels.

Download Full Size | PDF

We further notice that the effective bandgap EG + ε1,e + ε1,h also depends on the size Lx of a quantum dot due to the quantization effect, and the effective bandgap will increase with decreasing Lx. The size effect from such an Lx dependence is displayed in Fig. 6. From the upper-left panel of Fig. 6, we find that the peak of β0(ωsp) is enhanced as Lx is reduced. This phenomenon is connected to the increased localization of the SPP field at Lx = 170 Å as the SPP frequency approaches the saturation part of its dispersion. Moreover, the dip in β0(ωsp) is lifted somewhat uniformly at the same time due to a decreased n1e(t) from the enhanced Coulomb and phonon scattering at Lx = 170 Å. Here, β0(ωsp) is proportional to the population factor 1n1e(t)n1h(t), as can be seen from Eq. (63). Besides the slightly-reduced resonant peak strength of the scattered field for Lx = 170 Å (also resulting from the enhanced carrier scattering), |EtotEsp| keeps the same peak position, as shown in the upper-right panel of the figure. In this case, |EtotEsp| at the dot approaches a nonzero value at resonance, as can be seen from Eq. (55), and tends to zero rapidly away from resonance. Additionally, n2,h is reduced for Lx = 170 Å, as can be found from a comparison between the two lower panels of the figure. This is attributed to the reduced phonon absorption between two hole energy levels.

 figure: Fig. 6

Fig. 6 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as nℓ,e and holes nj,h (lower panels), as functions of Δh̄ωsp. Here, the results for three different sizes Lx = 210, 190 and 170 Å of a quantum dot are shown in the upper panels, along with a comparison of nℓ,e and holes nj,h for Lx = 210 and 170 Å in the lower panels.

Download Full Size | PDF

In Figs. 4 and 6, we vary the localization of an SPP field by changing the effective bandgap. Since the frequency of the surface plasmon (saturated dispersion part) is proportional to the factor of 1/1+εd, a smaller value of εd implies a higher surface-plasmon frequency or a reduced localization of the SPP field. We verify the change in the SPP localization by observing the upper two panels of Fig. 7, where the absorption peak, as well as the resonant scattered-field peak, become much stronger as εd is increased from 8 to 12 due to the reduction of saturated absorption for a lower field strength at the quantum dot. Furthermore, from the two lower panels of this figure we also observe, via the jumps in the population curves, an enhanced Rabi-split energy gap in the electron dressed states as εd is reduced from 12 to 10 due to the enhanced field strength at the quantum dot.

 figure: Fig. 7

Fig. 7 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as nℓ,e and holes nj,h (lower panels), as functions of Δh̄ωsp. Here, the results for three dielectric constants of a cladding layer, with εd = 8, 10 and 12, are displayed in the upper panels, along with a comparison of nℓ,e and holes nj,h for εd = 12 and 10 in the lower panels.

Download Full Size | PDF

In the presence of the localization of an SPP field, we can move a quantum dot closer to a metallic surface to gain a higher field at the quantum dot. The upper-left panel of Fig. 8 has elucidated this fact, in which a larger z0 corresponds to a weaker field, and then, a higher absorption peak due to the reduction of saturated absorption. This fact is also reflected in the upper-right panel of the figure, where a higher resonant scattered-field peak occurs for a larger value of z0. At z0 = 510 Å, a Rabi-split energy gap at resonance is clearly visible from the lower-left-panel of the figure for electron dressed states. Additionally, at z0 = 710 Å, by entering into a weak-coupling regime for a weaker field at the dot, we find sharpened resonant peaks in the energy-level occupations of electrons and holes, similar to the observation from the lower-left panel of Fig. 3.

 figure: Fig. 8

Fig. 8 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as nℓ,e and holes nj,h (lower panels), as functions of Δh̄ωsp. Here, the results for different vertical distances of a quantum dot from a metallic surface are displayed in the upper panels for z0 = 510, 610 and 710 Å, along with a comparison of nℓ,e and holes nj,h for z0 = 510 and 710 Å in the lower panels.

Download Full Size | PDF

3.2. Results for the dressed states of electrons

In the second part of the numerical calculations, besides the parameters given in the first sub-section, we have fixed Lx = 210 Å, h̄γ0 = 3 meV, z0 = 610 Å and εd = 12. Other parameters, including T, Esp and Δh̄ωsp, will be directly indicated in the figures. Additionally, Δh̄ωsp is given with respect to the energy gap at T = 300 K.

From the left panel of Fig. 9 we find a strong absorption (positive) peak and a weak gain (negative) peak for the probe-field absorption coefficient βabs(ωp) due to a quantum coherence effect from the electron states being dressed by an SPP field. In the strong-coupling regime, the dispersion of the quantum-dot e-h plasmas (dot-like branch) and SPPs (photon-like branch) form an anticrossing gap, where a higher-energy dot-like branch at a negative frequency detuning switches to a photon-like branch for a positive detuning. The positive peak is associated with the absorption of a probe-field photon by a quantum-dot e-h plasma, while the negative peak relates to the process with absorption of two photons from an SPP field and emission of one probe-field photon. The absorption peak is significantly reduced by saturation at Esp = 1000 kV/cm, and the gain peak is suppressed by a smaller Rabi-coupling frequency at Esp = 250 kV/cm (see the inset of the left panel). In addition, we observe from the right panel of Fig. 9 that two Rabi-splitting-induced side emission peaks for the spontaneous emission Pem(ω) become weaker and closer to the strong central peak as Esp is reduced (see the inset of the right panel). Moreover, the strength of the central peak due to the coherent conversion of an absorbed SPP-field photon to a spontaneously-emitted photon (non-linear optical behavior) is slightly reduced at Esp = 1000 kV/cm as a result of saturated absorption of the SPP field.

 figure: Fig. 9

Fig. 9 Probe-field absorption coefficient βabs(ωp) (left panel) and spontaneous emission of a quantum dot Pem(ω) (right panel) as functions of h̄ωp − (EG + ε1,e + ε1,h) and h̄ω − (EG + ε1,e + ε1,h), respectively, are presented. Here, comparisons of the results with three values of SPP-field amplitudes Esp = 250, 500 and 1000 kV/cm are given. The insets of both panels are enlarged views of the peaks.

Download Full Size | PDF

Figure 10 demonstrates the effect of frequency detuning Δh̄ωsp of an SPP field with respect to the bandgap of a quantum dot. The switching of the detuning from 10 meV to −10 meV reveals the corresponding spectral-position interchange between the absorption (dot-like branch) and the gain (photon-like branch) peaks for βabs(ωp) in the left panel of the figure. The Rabi oscillations between the first electron and hole energy levels are weakened with increasing |Δh̄ωsp|. At resonance with a zero detuning, both the absorption and gain peaks are suppressed by very strong Rabi oscillations. This detuning also shifts the emission peaks correspondingly because of the coherent conversion of an SPP-field photon to a spontaneously-emitted one, as can been seen from the right panel of this figure. Moreover, the central peak is weakened and the two side peaks are enlarged at resonance as a result of energy transfer to the side peaks by strong coupling and enhanced Rabi oscillations, respectively.

 figure: Fig. 10

Fig. 10 βabs(ωp) (left panel) and Pem(ω) (right panel) as functions of h̄ωp − (EG + ε1,e + ε1,h) and h̄ω − (EG + ε1,e + ε1,h), respectively, are displayed. Here, we show the comparisons with three values of SPP-field detunings Δh̄ωsp = 0 and ±10 meV. The label ×100 for zero SPP-field detuning indicates the result is multiplied by a factor of 100.

Download Full Size | PDF

Since the temperature affects the crystal bandgap energy EG, by changing the temperature we are able to scan the detuning Δh̄ωsp of the SPP field with a fixed SPP frequency h̄ωsp from negative to positive or vice versa. This leads to a spectral-position interchange between the absorption and gain peaks, similar to Fig. 10. The results in Fig. 11 prove such an expected feature by increasing T from 250 to 300 K in steps of 5 K. Technically, changing the temperature in the experiment is much easier than changing the tuning of a laser frequency over a large range. Here, the shift of the central peak in the right panel of the figure directly reflects the variation of the SPP-field detuning with T. Furthermore, the interchange between the dot-like and photon-like modes with T in the left panel can be regarded as direct evidence for the existence of an anticrossing energy gap resulting from a strongly-coupled e-h plasma and SPP field or coupled e-h plasmas and surface plasmons.

 figure: Fig. 11

Fig. 11 βabs(ωp) (left panel) and Pem(ω) (right panel) as functions of h̄ωp − (EG + ε1,e + ε1,h) and h̄ω − (EG + ε1,e + ε1,h), respectively, are shown. Here, a series of offset curves are displayed for various temperatures from T = 250 K (bottom black curves) to T = 300 K (top blue curves) in steps of 5 K.

Download Full Size | PDF

3.3. Time-resolved optical spectra

In our previously presented numerical results, we only showed steady-state dynamics of photo-excited e-h plasmas in a quantum dot by using a continuous SPP field, where the effects of both phonon scattering and e-h pair radiative recombination are combined with each other. Using a laser pulse to launch a pulsed SPP field, we are able to study the dynamics of phonon scattering (narrow pulse) as well as the dynamics of e-h pair radiative recombination (wide pulse), separately. Dynamically, phonon scattering becomes effective only after a characteristic time (around 1 ps), its effect can be seen from a significant increase of n2,e in our system. Figure 12 displays the results for β0(ωsp) (upper-left), |EtotEsp| (upper-right), n1,e (lower-left) and n2,e (lower-right) for various detection times τ0 in the presence of a narrow laser pulse (with pulse width Tp = 500 fs and peak value Esp = 500kV/cm) which is turned on at t = 0. We see from Fig. 12 that β0(ωsp) starts with a dip for the dressed state at resonance, then shifts to a single peak (at half-pulse width) due to a suppression of the photon-dressing effect. It eventually becomes a single peak plus a shifted dip after the pulse has passed due to formation of resonant peaks in n1,e and n1,h. Correspondingly, |EtotEsp| starts by showing a non-resonant behavior with a relatively large magnitude, then shifts to a quasi-resonant behavior, and finally looks like suppressed resonant behavior with a peak at and dips on both sides of Δh̄ωsp = 0. The resonant build up of n1,e after τ0 ≥ 500 fs can also be verified from this figure, which is accompanied by the start of significant phonon absorption after τ0 ≥ 1 ps.

 figure: Fig. 12

Fig. 12 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as n1,e and holes n2,e (lower panels), as functions of Δh̄ωsp at different detection times τ0 for a Gaussian-shape laser pulse with pulse width Tp = 500 fs. Here, Esp = 500 kV/cm is taken, and the other parameters are the same as those in Fig. 3. The labels ×10 and ×105 in the upper-right panel indicate that the results are multiplied by factors of 10 and 105, respectively.

Download Full Size | PDF

Technically, detecting dynamics of photo-excited e-h plasmas by using another time-delayed weak probe field is much more feasible, as shown in Fig. 13. From the left panel of this figure, we find that βabs(ωp) starts with a pair of positive absorption and negative gain peaks due to a very strong photon dressing effect for the delay times τd = 60 and 120 fs. This is changed to a strong absorption peak plus a very weak gain peak at τd = 240 fs. At the end, βabs(ωp) becomes independent of τd, indicating that a linear optical-response regime has been reached. On the other hand, from the right panel of this figure, we see that the central peak of Pem(ω) is gradually built up with increasing τd due to enhanced n1,e and n1,h around resonance, while two side peaks become weakened and disappear at the same time due to weakened Rabi oscillations. Interestingly, we also find that the central peak of Pem(ω) slightly decreases at τd = 1 ps, which agrees with the observed start of significant phonon absorption seen in the lower-left panel of Fig. 12.

 figure: Fig. 13

Fig. 13 βabs(ωp) (left panel) and Pem(ω) (right panel) as functions of h̄ωp − (EG + ε1,e + ε1,h) and h̄ω − (EG + ε1,e + ε1,h), respectively, at different delay times τd for a Gaussian-shape laser pulse with pulse width Tp = 500 fs. Here, Esp = 500 kV/cm and Δh̄ωsp = 5 meV are chosen, and the other parameters are the same as those in Fig. 3.

Download Full Size | PDF

In order to explore the dynamics of e-h pair radiative recombination in our system, a wide pulse with a full-pulse width around 300 ps is required, as displayed in Fig. 14. From the upper-middle panel of this figure, we find that β0(ωsp) starts with a resonant dip due to a strong photon dressing effect, then shifts to a sole peak at Δh̄ωsp = 0 as τ0 ≥ 400 ps where a steady state is almost reached in the linear-response regime. Accordingly, the level populations n1,e and n2,e in the lower two panels show a transition from an initial non-resonant behavior to a final resonant behavior. This is accompanied by dramatically reduced level populations due to the start of a radiative recombination process for photo-excited e-h pairs.

 figure: Fig. 14

Fig. 14 β0(ωsp) (upper-middle panel), n1,e and holes n2,e (lower two panels), as functions of Δh̄ωsp at different detection times τ0 for a wide laser pulse with pulse width Tp = 300 ps. Here, Esp = 500 kV/cm is taken, and the other parameters are the same as those in Fig. 3. The labels ×0.05 in the upper-middle panel indicate that the results in the upper panel for τ0 = 60 and 120 ps are multiplied by a factor of 0.05.

Download Full Size | PDF

Recombination dynamics for e-h plasmas can also be demonstrated clearly by the time-delayed probe-field absorption as well as by the time-resolved spontaneous emission, as shown in Fig. 15. As presented in the left panel of this figure, we find that the initial weak absorption and gain peaks (see the inset) in βabs(ωp) occur at τd = 200 ps and are replaced by a strong single absorption peak due to a suppressed photon dressing effect and phase-space blocking.

 figure: Fig. 15

Fig. 15 βabs(ωp) (left panel) and Pem(ω) (right panel) as functions of h̄ωp − (EG + ε1,e + ε1,h) and h̄ω − (EG + ε1,e + ε1,h), respectively, at different delay times τd for a Gaussian-shape laser pulse with pulse width Tp = 300 ps. Here, Esp = 500 kV/cm and Δh̄ωsp = 5 meV are assumed, and the other parameters are the same as those in Fig. 3. The inset in the upper-left panel shows an enlarged view for the gain and absorption peaks for small delay times. The labels ×10 for τd = 50 and 100 ps in the left panel, as well as the label ×0.02 in the right panel, indicate that the results are multiplied by a factor of 10 and a factor of 0.02, respectively.

Download Full Size | PDF

On the other hand, from the right panel of the same figure, we see that the initial central peak in Pem(ω) is increased very rapidly due to accumulation of photo-excited e-h pairs and accompanied by the reduction of two side peaks resulting from the weakened Rabi oscillations. Importantly, the very-strong central peak in Pem(ω) is significantly reduced at τd = 200 ps, indicating the start of a radiative-recombination process for photo-excited e-h plasmas. This recombination process is continuously enhanced with the increasing delay time τd and suppresses the central peak in Pem(ω) after τd ≥ 400 ps due to draining out the photo-generated electrons and holes at the same time.

4. Conclusions and remarks

In conclusion, we have demonstrated the possibility of using a SPP field to control the optical gain and absorption of another passing light beam due to their strong nonlinear field coupling mediated by electrons in the quantum dot. We have also predicted the coherent conversion of a surface-plasmon-field photon to a spontaneously-emitted free-space photon, which is simultaneously accompanied by another pair of blue- and red-shifted photons.

Although we studied only the coupling of a SPP field to a single quantum dot in this paper for the simplest case, our formalism can be generalized easily to include many quantum dots. The numerically-demonstrated unique control of the effective photon-photon coupling by the quantum dot can be used for constructing an optical transistor, where the ‘gate’ photon controls the intensity of its ‘source’ light beam. These optical transistors are very useful for speeding up and improving the performance of fiber-optic communication networks, as well as for constructing quantum information and developing optical digital computers.

Furthermore, instead of a resonant coupling to the lowest pair of electron-hole energy levels, we may select the surface-plasmon frequency for resonant coupling to the higher pair of electron-hole levels. In such a case, the optical pumping from the intense surface-plasmon near-field could create a population inversion with respect to the ground pair of electron-hole levels by emitting thermal phonons, leading to a possible lasing action if the optical gain can overcome the metal loss for the surface plasmons. Such a surface-plasmon quantum-dot laser would have a beam size as small as a few nanometers beyond the optical diffraction limit, and it is expected to be very useful for spatially-selective illumination of individual molecules or neuron cells in low-temperature photo-excited chemical reactions or optogenetics and neuroscience.

A. Results Related to the Electronic States of a Quantum Dot

We have employed a box-type potential with hard walls to model a quantum dot, which is given by

V(r)={0,0xiLifori=1,2,3,otherwise,
where the position vector r = (x1, x2, x3), and L1, L2 and L3 are the widths of the potential in the x1, x2 and x3 directions, respectively. The Schrödinger equation for a single electron or hole in a quantum dot is written as
h¯22m*[2x12+2x22+2x32+V(r)]ψ(r)=εψ(r),
where the effective mass m* is me* for electrons or mh* for holes. The eigenstate wave-function associated with Eq. (74) is found to be
ψn1,n2,n3(r)=2L1sin[(n1πL1)x1]2L2sin[(n2πL2)x2]2L3sin[(n3πL3)x3],
which is same for both electrons and holes, and the eigenstate energy associated with Eq. (74) is
εn1,n2,n3=h¯22m*[(n1πL1)2+(n2πL2)2+(n3πL3)2],
where the quantum numbers n1, n2, n3 = 1, 2, ···.

By using the calculated bare energy levels in Eq. (76), the dressed electron ( λαe) and hole ( λαh) energy levels under the rotating wave approximation take the forms [2]

λαe(ω|t)=λαh(ω|t)={12(h¯ω+[G(T)+εαe+εαhh¯ω]2+4|α,αeh(t)|2)ifh¯ωG(T)+εαe+εαh12(h¯ω[G(T)+εαe+εαhh¯ω]2+4|α,αeh(t)|2)ifh¯ωG(T)+εαe+εαh,
where the composite index α = {n1, n2, n3}. Moreover, we obtain the energy levels of dressed electrons ε¯αe(ω|t)=λαe(ω|t)+(εαeεαh)/2 and ε¯e(ω|t)=εe+G(T)/2 for α. Similarly, we obtain the energy levels of dressed holes ε¯αh(ω|t)=λαh(ω|t)+(εαhεαe)/2 and ε¯je(ω|t)=εjh+G(T)/2 for jα.

Based on the calculated wave-functions in Eq. (75), the form factors introduced in Eqs. (11) and (12) can be obtained from

n1,n2,n3;n1,n2,n3e(q)=n1,n2,n3;n1,n2,n3h(q)=𝒬n1,n1(1)(q1)𝒬n2,n2(2)(q2)𝒬n3,n3(3)(q3),
where the wave vector q = (q1, q2, q3) and we have introduced the following notation for j = 1, 2, 3
𝒬nj,njj(qj)=(2Lj)0Ljdxjeiqjxjsin[(njπLj)xj]sin[(njπLj)xj].
Moreover, the overlap of the electron and hole wave-functions in this model can be easily calculated as
d3rψn1,n2,n3e(r)ψn1,n2,n3h(r)=δn1,n1δn2,n2δn3,n3.
The interband dipole moment dc,v = dc,v êd at the isotropic Γ-point, which is defined in Eq. (32), can be calculated according to the Kane approximation [35, 36]
dc,v=e2h¯22m0G(T)(m0me*1).
Furthermore, the direction of the dipole moment êd is determined by the quantum-dot energy levels in resonance with the photon energy h̄ω.

Acknowledgments

DH would like to thank the Air Force Office of Scientific Research (AFOSR) for support. GG was supported by contract # FA 9453-13-1-0291 of AFRL.

References and links

1. O. Firstenberg, T. Peyronel, Q.-Y. Liang, A. V. Gorshkov, M. D. Lukin, and V. Vuletić, “Attractive photons in a quantum nonlinear medium,” Nat. 502, 71–75 (2013). [CrossRef]  

2. S. Schmitt-Rink, D. S. Chemla, and H. Haug, “Nonequilibrium theory of the optical Stark effect and spectral hole burning in semiconductors,” Phys. Rev. B 37, 941–955 (1988). [CrossRef]  

3. I. Fushman, D. Englund, A. Faraon, N. Stoltz, P. Petroff, and J. Vučković, “Controlled phase shifts with a single quantum dot,” Sci. 320, 769–772 (2008). [CrossRef]  

4. W. Chen, K. M. Beck, R. Bücker, M. Gullans, M. D. Lukin, H. Tanji-Suzuki, and V. Vuletić, “All-optical switch and transistor gated by one stored photon,” Sci. 341, 768–770 (2013). [CrossRef]  

5. G. Gumbs and D. H. Huang, Properties of Interacting Low-Dimensional Systems (Wiley-VCH Verlag GmbH & Co. kGaA, 2011), Chaps. 4, 5. [CrossRef]  

6. D. H. Huang, G. Gumbs, and S.-Y. Lin, “Self-consistent theory for near-field distribution and spectrum with quantum wires and a conductive grating in terahertz regime,” J. Appl. Phys. 105, 093715 (2009). [CrossRef]  

7. D. Dini, R. Köhler, A. Tredicucci, G. Biasiol, and L. Sorba, “Microcavity polariton splitting of intersubband transitions,” Phys. Rev. Lett. 90, 116401 (2003). [CrossRef]   [PubMed]  

8. Y. Todorov, A. M. Andrews, I. Sagnes, R. Colombelli, P. Klang, G. Strasser, and C. Sirtori, “Strong light-matter coupling in subwavelength metal-dielectric microcavities at terahertz frequencies,” Phys. Rev. Lett. 102, 186402 (2009). [CrossRef]   [PubMed]  

9. Y. Todorov, A. M. Andrews, R. Colombelli, S. De Liberato, C. Ciuti, P. Klang, G. Strasser, and C. Sirtori, “Ultrastrong light-matter coupling regime with polariton dots,” Phys. Rev. Lett. 105, 196402 (2010). [CrossRef]  

10. H. Deng, H. Haug, and Y. Yamamoto, “Exciton-polariton Bose-Einstein condensation,” Rev. Mod. Phys. 82, 1489–1537 (2010). [CrossRef]  

11. S. Christopoulos, G. B. H. von Högersthal, A. J. D. Grundy, P. G. Lagoudakis, A.V. Kavokin, J. J. Baumberg, G. Christmann, R. Butté, E. Feltin, J.-F. Carlin, and N. Grandjean, “Room-temperature polariton lasing in semiconductor microcavities,” Phys. Rev. Lett. 98, 126405 (2007). [CrossRef]   [PubMed]  

12. S. I. Tsintzos, N. T. Pelekanos, G. Konstantinidis, Z. Hatzopoulos, and P. G. Savvidis, “A GaAs polariton light-emitting diode operating near room temperature,” Nat. Lett. 453, 372–375 (2008). [CrossRef]  

13. P. Bhattacharya, B. Xiao, A. Das, S. Bhowmick, and J. Heo, “Solid state electrically injected exciton-polariton laser,” Phys. Rev. Lett. 110, 206403 (2013). [CrossRef]   [PubMed]  

14. C. Schneider, A. Rahimi-Iman, N. Y. Kim, J. Fischer, I. G. Savenko, M. Amthor, M. Lermer, A. Wolf, L. Worschech, V. D. Kulakovskii, I. A. Shelykh, M. Kamp, S. Reitzenstein, A. Forchel, Y. Yamamoto, and S. Höfling, “An electrically pumped polariton laser,” Nat. 497, 348–352 (2013). [CrossRef]  

15. F. Jahnke, M. Kira, and S. W. Koch, “Linear and nonlinear optical properties of excitons in semiconductor quantum wells and microcavities,” Z. Phys. B 104, 559–572 (1997). [CrossRef]  

16. D. H. Huang, M. M. Easter, G. Gumbs, A. A. Maradudin, S.-Y. Lin, D. A. Cardimona, and X. Zhang, “Resonant scattering of surface plasmon polaritons by dressed quantum dots,” Appl. Phys. Lett. 104, 251103 (2014). [CrossRef]  

17. Y. H. Cao, J. Xie, Y. M. Liu, and Z. Y. Liu, “Modeling and optimization of photonic crystal devices based on transformation optics method,” Opt. Express 22, 2725–2734 (2014). [CrossRef]   [PubMed]  

18. X. Yang, J. Yao, J. Rho, X. Yin, and X. Zhang, “Experimental realization of three-dimensional indefinite cavities at the nanoscale with anomalous scaling laws,” Nat. Photonics 6, 450–454 (2012). [CrossRef]  

19. M. Lindberg, Y. Z. Hu, R. Binder, and S. W. Koch, “χ(3) formalism in optically excited semiconductors and its applications in four-wave-mixing spectroscopy,” Phys. Rev. B 50, 18060–18072 (1994). [CrossRef]  

20. D. H. Huang and P. M. Alsing, “Many-body effects on optical carrier cooling in intrinsic semiconductors at low lattice temperatures,” Phys. Rev. B 78, 035206 (2008). [CrossRef]  

21. H. M. Gibbs, G. Khitrova, and S. W. Koch, “Exciton-polariton light-semiconductor coupling effects,” Nat. Photonics 5, 275–282 (2011). [CrossRef]  

22. T. Yoshie, A. Scherer, J. Hendrickson, G. Khitrova, H. M. Gibbs, G. Rupper, C. Ell, O. B. Shchekin, and D. G. Deppe, “Vacuum Rabi splitting with a single quantum dot in a photonic crystal nanocavity,” Nat. 432, 200–203 (2004). [CrossRef]  

23. R. F. Oulton, V. J. Sorger, T. Zentgraf, R.-M. Ma, C. Gladden, L. Dai, G. Bartal, and X. Zhang, “Plasmon lasers at deep subwavelength scale,” Nat. Lett. 461, 629–632 (2009). [CrossRef]  

24. F. Alpeggiani, S. D’Agostino, and L. C. Andreani, “Surface plasmons and strong light-matter coupling in metallic nanoshells,” Phys. Rev. B 86, 035421 (2012). [CrossRef]  

25. A. A. Maradudin and D. L. Mills, “Scattering and absorption of electromagnetic radiation by a semi-infinite medium in the presence of surface roughness,” Phys. Rev. B 11, 1392–1415 (1975). [CrossRef]  

26. M. G. Cottam and A. A. Maradudin, “Surface linear response functions,” in Surface Excitations, eds. V. M. Agranovich and R. Loudon, eds. (North-Holland, Amsterdam, 1984), pp. 1–194.

27. R. V. Shenoi, J. Bur, D. H. Huang, and S.-Y. Lin, “Extraordinary plasmon-QD interaction for enhanced infrared application,” Proc. SPIE 8632, Photonic and Phononic Properties of Engineered Nanostructures III, 86321L (2013). [CrossRef]  

28. R. V. Shenoi, S.-Y. Lin, S. Krishna, and D. H. Huang, “Order-of-magnitude enhancement of intersubband photo-response in a plasmonic-quantum dot system,” Opt. Lett. 39, 4454–4457 (2014). [CrossRef]   [PubMed]  

29. M. S. Tame, K. R. McEnery, S. K. Özdemir, J. Lee, S. A. Maier, and M. S. Kim, “Quantum plasmonics,” Nat. Phys. 9, 329–340 (2013). [CrossRef]  

30. F. Rossi and T. Kuhn, “Theory of ultrafast phenomena in photoexcited semiconductors,” Rev. Mod. Phys. 74, 895–950 (2002). [CrossRef]  

31. V. M. Axt and T. Kuhn, “Femtosecond spectroscopy in semiconductors: a key to coherences, correlations and quantum kinetics,” Rep. Prog. Phys. 67, 433–512 (2004). [CrossRef]  

32. M. Kira and S. W. Koch, “Many-body correlations and excitonic effects in semiconductor spectroscopy,” Prog. Quant. Electr. 30, 155–296 (2006). [CrossRef]  

33. D. H. Huang, P. M. Alsing, T. Apostolova, and D. A. Cardimona, “Coupled energy-drift and force-balance equations for high-field hot-carrier transport,” Phys. Rev. B 71, 195205 (2005). [CrossRef]  

34. D. H. Huang and D. A. Cardimona, “Intersubband laser coupled three-level asymmetric quantum wells: new dynamics of quantum interference,” J. Opt. Soc. Am. B 15, 1578—1584 (1998). [CrossRef]  

35. E. O. Kane, “Band structure of indium antimonide,” J. Physi. Chem. Solids 1, 249–261 (1957). [CrossRef]  

36. U. Bockelmann and G. Bastard, “Interband absorption in quantum wires. I. Zero-magnetic-field case,” Phys. Rev. B 45, 1688–1699 (1992). [CrossRef]  

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 (15)

Fig. 1
Fig. 1 Schematic illustration for a semi-infinite metal and a quantum dot above its surface at x3 = 0. Here, the surface-plasmon polariton (SPP) is locally excited by incident light with the help of a surface grating. The propagating SPP field further excites e-h pairs (plasmas) in the adjacent quantum dot. As a result, the optical-polarization field of the photo-excited e-h plasma is strongly coupled to the propagating SPP field to form split plasma-SPP modes with an anticrossing gap. Also, a probe-field is used for studying the photon dressing effect.
Fig. 2
Fig. 2 Schematic illustration for a system incorporating the generation of quantum-dot excitons by a SPP field with frequency ωsp and probed by a plane-wave field Ep with frequency ωp. Here, G is the energy bandgap of the host semiconductor, ε e and ε j h stand for the energy levels of electrons and holes, respectively, with , j = 1, 2, ···. In addition, ωem represents the frequency of spontaneous emission from photo-excited excitons, and the ground states of electrons and holes are coupled to their first excited states by lattice phonons at finite temperatures.
Fig. 3
Fig. 3 Optical absorption coefficients β0(ωsp) (upper-left panel) and scattering field |EtotEsp| at the quantum dot (upper-right panel), as well as the energy-level occupations for electrons nℓ,e and holes nj,h (lower panels) with , j = 1, 2, as functions of the frequency detuning Δh̄ωsph̄ωsp − (EG +ε1,e +ε1,h). Here, the results for various amplitudes Esp of an SPP field with frequency ωsp are presented in the upper panels, along with a comparison of the energy-level occupations for Esp = 25 and 125 kV/cm in the lower panels. The label ×0.1 in the upper-left panel indicates that the result is multiplied by a factor of 0.1.
Fig. 4
Fig. 4 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as nℓ,e and holes nj,h (lower panels), as functions of Δh̄ωsp. Here, the results for three different temperatures T = 300, 175 and 77 K are displayed in the upper panels, along with a comparison of nℓ,e and holes nj,h for T = 300 and 77 K in the lower panels.
Fig. 5
Fig. 5 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as nℓ,e and holes nj,h (lower panels), as functions of Δh̄ωsp. Here, the results with different energy-level broadening h̄γ0 = 3, 5 and 7 meV are shown in the upper panels, along with a comparison of nℓ,e and holes nj,h for h̄γ0 = 3 and 7 meV in the lower panels.
Fig. 6
Fig. 6 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as nℓ,e and holes nj,h (lower panels), as functions of Δh̄ωsp. Here, the results for three different sizes Lx = 210, 190 and 170 Å of a quantum dot are shown in the upper panels, along with a comparison of nℓ,e and holes nj,h for Lx = 210 and 170 Å in the lower panels.
Fig. 7
Fig. 7 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as nℓ,e and holes nj,h (lower panels), as functions of Δh̄ωsp. Here, the results for three dielectric constants of a cladding layer, with εd = 8, 10 and 12, are displayed in the upper panels, along with a comparison of nℓ,e and holes nj,h for εd = 12 and 10 in the lower panels.
Fig. 8
Fig. 8 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as nℓ,e and holes nj,h (lower panels), as functions of Δh̄ωsp. Here, the results for different vertical distances of a quantum dot from a metallic surface are displayed in the upper panels for z0 = 510, 610 and 710 Å, along with a comparison of nℓ,e and holes nj,h for z0 = 510 and 710 Å in the lower panels.
Fig. 9
Fig. 9 Probe-field absorption coefficient βabs(ωp) (left panel) and spontaneous emission of a quantum dot Pem(ω) (right panel) as functions of h̄ωp − (EG + ε1,e + ε1,h) and h̄ω − (EG + ε1,e + ε1,h), respectively, are presented. Here, comparisons of the results with three values of SPP-field amplitudes Esp = 250, 500 and 1000 kV/cm are given. The insets of both panels are enlarged views of the peaks.
Fig. 10
Fig. 10 βabs(ωp) (left panel) and Pem(ω) (right panel) as functions of h̄ωp − (EG + ε1,e + ε1,h) and h̄ω − (EG + ε1,e + ε1,h), respectively, are displayed. Here, we show the comparisons with three values of SPP-field detunings Δh̄ωsp = 0 and ±10 meV. The label ×100 for zero SPP-field detuning indicates the result is multiplied by a factor of 100.
Fig. 11
Fig. 11 βabs(ωp) (left panel) and Pem(ω) (right panel) as functions of h̄ωp − (EG + ε1,e + ε1,h) and h̄ω − (EG + ε1,e + ε1,h), respectively, are shown. Here, a series of offset curves are displayed for various temperatures from T = 250 K (bottom black curves) to T = 300 K (top blue curves) in steps of 5 K.
Fig. 12
Fig. 12 β0(ωsp) (upper-left panel) and |EtotEsp| (upper-right panel), as well as n1,e and holes n2,e (lower panels), as functions of Δh̄ωsp at different detection times τ0 for a Gaussian-shape laser pulse with pulse width Tp = 500 fs. Here, Esp = 500 kV/cm is taken, and the other parameters are the same as those in Fig. 3. The labels ×10 and ×105 in the upper-right panel indicate that the results are multiplied by factors of 10 and 105, respectively.
Fig. 13
Fig. 13 βabs(ωp) (left panel) and Pem(ω) (right panel) as functions of h̄ωp − (EG + ε1,e + ε1,h) and h̄ω − (EG + ε1,e + ε1,h), respectively, at different delay times τd for a Gaussian-shape laser pulse with pulse width Tp = 500 fs. Here, Esp = 500 kV/cm and Δh̄ωsp = 5 meV are chosen, and the other parameters are the same as those in Fig. 3.
Fig. 14
Fig. 14 β0(ωsp) (upper-middle panel), n1,e and holes n2,e (lower two panels), as functions of Δh̄ωsp at different detection times τ0 for a wide laser pulse with pulse width Tp = 300 ps. Here, Esp = 500 kV/cm is taken, and the other parameters are the same as those in Fig. 3. The labels ×0.05 in the upper-middle panel indicate that the results in the upper panel for τ0 = 60 and 120 ps are multiplied by a factor of 0.05.
Fig. 15
Fig. 15 βabs(ωp) (left panel) and Pem(ω) (right panel) as functions of h̄ωp − (EG + ε1,e + ε1,h) and h̄ω − (EG + ε1,e + ε1,h), respectively, at different delay times τd for a Gaussian-shape laser pulse with pulse width Tp = 300 ps. Here, Esp = 500 kV/cm and Δh̄ωsp = 5 meV are assumed, and the other parameters are the same as those in Fig. 3. The inset in the upper-left panel shows an enlarged view for the gain and absorption peaks for small delay times. The labels ×10 for τd = 50 and 100 ps in the left panel, as well as the label ×0.02 in the right panel, indicate that the results are multiplied by a factor of 10 and a factor of 0.02, respectively.

Equations (81)

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

× × E ( r ; ω ) ε b ( x 3 ; ω ) ω 2 c 2 E ( r ; ω ) = ω 2 ε 0 c 2 𝒫 loc ( r ; ω ) ,
ε b ( x 3 ; ω ) = { ε d , for x 3 > 0 ε M ( ω ) , for x 3 < 0 .
μ [ ε b ( x 3 ; ω ) ω 2 c 2 δ λ μ 2 x λ x μ + δ λ μ r 2 ] 𝒢 μ ν ( r , r ; ω ) = δ λ ν δ ( r r ) ,
E μ ( r ; ω ) = E μ ( 0 ) ( r ; ω ) ω 2 ε 0 c 2 ν d 3 r 𝒢 μ ν ( r , r ; ω ) 𝒫 ν loc ( r ; ω ) ,
ν [ ε b ( x 3 ; ω ) ω 2 c 2 δ μ ν 2 x μ x ν + δ μ ν r 2 ] E ν ( 0 ) ( r ; ω ) = 0 ,
𝒢 μ ν ( r , r ; ω ) = d 2 k ( 2 π ) 2 e i k ( r r ) g μ ν ( k , ω | x 3 , x 3 ) ,
[ ε b ω 2 c 2 k 2 2 + d 2 d x 3 2 k 1 k 2 i k 1 d d x 3 k 1 k 2 ε b ω 2 c 2 k 1 2 + d 2 d x 3 2 i k 2 d d x 3 i k 1 d d x 3 i k 2 d d x 3 ε b ω 2 c 2 k 2 ] [ g 11 g 12 g 13 g 21 g 22 g 23 g 31 g 32 g 33 ] = δ ( x 3 x 3 ) [ 1 0 0 0 1 0 0 0 1 ] .
f μ ν ( k , ω | x 3 , x 3 ) = μ , ν 𝒮 μ μ ( k ) 𝒮 ν ν ( k ) g μ ν ( k , ω | x 3 , x 3 ) ,
𝒮 ( k ) = 1 k [ k 1 k 2 0 k 2 k 1 0 0 0 k ] ,
[ ε b ω 2 c 2 + d 2 d x 3 2 0 i k d d x 3 0 ε b ω 2 c 2 k 2 + d 2 d x 3 2 0 i k d d x 3 0 ε b ω 2 c 2 k 2 ] [ f 11 f 12 f 13 f 21 f 22 f 23 f 31 f 32 f 33 ] = δ ( x 3 x 3 ) [ 1 0 0 0 1 0 0 0 1 ] .
f 22 ( k , ω | x 3 , x 3 ) = { ( i 2 p ) 2 p p d + p e i p d x 3 i p x 3 x 3 > 0 , x 3 < 0 ( i 2 p ) [ e i p | x 3 x 3 | p d p p d + p e i p ( x 3 + x 3 ) ] , x 3 < 0 , x 3 < 0 ( i 2 p d ) [ e i p d | x 3 x 3 | + p d p p d + p e i p d ( x 3 + x 3 ) ] , x 3 > 0 , x 3 > 0 ( i 2 p d ) 2 p d p d + p e i p ( x 3 x 3 ) , x 3 < 0 , x 3 > 0 ,
f 13 ( k , ω | x 3 , x 3 ) = { i k c 2 2 ε M ( ω ) ω 2 [ 2 ε M ( ω ) p d ε M ( ω ) p d + ε d p ] e i p d x 3 i p x 3 , x 3 > 0 , x 3 < 0 i k c 2 2 ε M ( ω ) ω 2 [ e i p | x 3 x 3 | sgn ( x 3 x 3 ) + ε M ( ω ) p d ε d p ε M ( ω ) p d + ε d p e i p ( x 3 + x 3 ) ] , x 3 < 0 , x 3 < 0 i k c 2 2 ε d ω 2 [ e i p d | x 3 x 3 | sgn ( x 3 x 3 ) + ε M ( p d ) p d ε d p ε M ( ω ) p d + ε d p e i p d ( x 3 + x 3 ) ] , x 3 > 0 , x 3 > 0 i k c 2 2 ε d ω 2 [ 2 ε d p ε M ( ω ) p d + ε d p ] e i p x 3 + i p d x 3 , x 3 < 0 , x 3 > 0
f 33 ( k , ω | x 3 , x 3 ) = { i k 2 c 2 ω 2 [ 1 ε M ( ω ) p d + ε d p ] e i p d x 3 i p x 3 , x 3 > 0 , x 3 < 0 c 2 ε M ( ω ) ω 2 δ ( x 3 x 3 ) i k 2 c 2 2 p ε M ( ω ) ω 2 [ e i p | x 3 x 3 | ε M ( ω ) p d ε d p ε M ( ω ) p d + ε d p e i p ( x 3 + x 3 ) ] , x 3 < 0 , x 3 < 0 c 2 ε d ω 2 δ ( x 3 x 3 ) i k 2 c 2 2 p d ε d ω 2 [ e i p d | x 3 x 3 | + ε M ( ω ) p d ε d p ε M ( ω ) p d + ε d p e i p d ( x 3 + x 3 ) ] , x 3 > 0 , x 3 > 0 i k 2 c 2 ω 2 [ 1 ε M ( ω ) p d + ε d p ] e i p x 3 + i p d x 3 , x 3 < 0 , x 3 > 0
f 11 ( k , ω | x 3 , x 3 ) = { i p d p c 2 ω 2 [ 1 ε M ( ω ) p d + ε d p ] e i p d x 3 i p x 3 , x 3 > 0 , x 3 < 0 i p c 2 2 ε M ( ω ) ω 2 [ e i p | x 3 x 3 | + ε M ( ω ) p d ε d p ε M ( ω ) p d + ε d p e i p ( x 3 + x 3 ) ] , x 3 < 0 , x 3 < 0 i p d c 2 2 ε d ω 2 [ e i p d | x 3 x 3 | ε M ( ω ) p d ε d p ε M ( ω ) p d + ε d p e i p d ( x 3 + x 3 ) ] , x 3 > 0 , x 3 > 0 i p d c 2 2 ε d ω 2 [ 2 ε d p ε M ( ω ) p d + ε d p ] e i p x 3 + i p d x 3 , x 3 < 0 , x 3 > 0
f 31 ( k , ω | x 3 , x 3 ) = { i k c 2 ω 2 [ p ε M ( ω ) p d + ε d p ] e i p d x 3 i p x 3 , x 3 > 0 , x 3 < 0 i k c 2 2 ε M ( ω ) ω 2 [ e i p | x 3 x 3 | sgn ( x 3 x 3 ) ε M ( ω ) p d ε d p ε M ( ω ) p d + ε d p e i p ( x 3 + x 3 ) ] , x 3 < 0 , x 3 < 0 i k c 2 2 ε d ω 2 [ e i p d | x 3 x 3 | sgn ( x 3 x 3 ) ε M ( ω ) p d ε d p ε M ( ω ) p d + ε d p e i p d ( x 3 + x 3 ) ] , x 3 > 0 , x 3 > 0 i k c 2 ω 2 [ p d ε M ( ω ) p d + ε d p ] e i p x 3 + i p d x 3 , x 3 < 0 , x 3 > 0
p d ( k , ω ) = ε d ω 2 c 2 k 2 ,
p ( k , ω ) = ε M ( ω ) ω 2 c 2 k 2 ,
g μ ν ( k , ω | x 3 , x 3 ) = μ , ν f μ ν ( k , ω | x 3 , x 3 ) 𝒮 μ μ ( k ) 𝒮 ν ν ( k ) ,
d n e d t = 2 h ¯ j Im [ ( Y j ) * ( , j eh Y j V , j ; j , eh ) ] + n e t | rel δ , 1 sp n 1 e n 1 h ,
d n j h d t = 2 h ¯ Im [ ( Y j ) * ( , j eh Y j V , j ; j , eh ) ] + n j h t | rel δ j , 1 sp n 1 e n 1 h ,
N e ( t ) = 2 n e ( t ) = 2 j n j h ( t ) = N h ( t ) ,
i h ¯ d d t Y j = [ ε ¯ e ( ω ) + ε ¯ j h ( ω ) h ¯ ( ω + i γ 0 ) ] Y j + ( 1 n e n j h ) ( , j eh Y j V , j ; j , eh ) + Y j [ j 1 n j 1 h ( V j , j 1 ; j 1 , j hh V j , j 1 ; j , j 1 hh ) 1 n 1 e V 1 , j ; j , 1 eh ] + Y j [ 1 n 1 e ( V , 1 ; 1 , ee V , 1 ; , 1 ee ) j 1 n j 1 h V , j 1 ; j 1 , eh ] ,
Y j ( t | ω ) = [ 1 n e ( t ) n j h ( t ) h ¯ ( ω + i γ 0 ) h ¯ Ω ¯ , j eh ( ω | t ) ] , j eh ( t ) ,
h ¯ Ω ¯ , j eh ( ω | t ) = ε ¯ e ( ω | t ) + ε ¯ j h ( ω | t ) V , j ; j , e h + 1 n 1 e ( t ) ( V , 1 ; 1 , ee V , 1 ; , 1 ee ) + j 1 n j 1 h ( t ) ( V j , j 1 ; j 1 , j hh V j , j 1 ; j , j 1 hh ) 1 n 1 e ( t ) V 1 , j ; j , 1 eh j 1 j n j 1 h ( t ) V , j 1 ; j 1 , eh .
V 1 , 2 ; 3 , 4 ee = d 3 r d 3 r [ ψ 1 e ( r ) ] * [ ψ 2 e ( r ) ] * e 2 4 π ε 0 ε b | r r | ψ 3 e ( r ) ψ 4 e ( r ) = e 2 8 π 2 ε 0 ε b d 2 q 1 , 4 e ( q ) 2 , 3 e ( q ) ( 1 q + q s ) = ( V 1 , 2 ; 3 , 4 ee ) * ,
V j 1 , j 2 ; j 3 , j 4 hh = d 3 r d 3 r [ ψ j 1 h ( r ) ] * [ ψ j 2 h ( r ) ] * e 2 4 π ε 0 ε b | r r | ψ j 3 h ( r ) ψ j 4 h ( r ) = e 2 8 π 2 ε 0 ε b d 2 q j 1 , j 4 h ( q ) j 2 , j 3 h ( q ) ( 1 q + q s ) = ( V j 1 , j 2 ; j 3 , j 4 hh ) * ,
V , j ; j , eh = d 3 r d 3 r [ ψ e ( r ) ] * [ ψ j h ( r ) ] * e 2 4 π ε 0 ε b | r r | ψ j h ( r ) ψ e ( r ) = e 2 8 π 2 ε 0 ε b d 2 q , e ( q ) j , j h ( q ) ( 1 q + q s ) = ( V , j ; j , eh ) * ,
q s ( t ) = e 2 4 ε 0 ε b 𝒮 k B T [ N e ( t ) + N h ( t ) ] .
, e ( q ) = e q 0 d 2 r [ ψ e ( r ) ] * e i q r ψ e ( r ) = [ , e ( q ) ] * ,
j , j h ( q ) = e q 0 d 2 r [ ψ j h ( r ) ] * e i q r ψ j h ( r ) = [ j , j h ( q ) ] * ,
, j eh ( t ) = δ , 1 δ j , 1 θ ( t ) [ E , j eh ( ω ) d c , v ] ,
d c , v = d 3 r [ u c ( r ) ] * r u v ( r ) = d c , v * ,
E , j eh ( ω ) = d 3 r [ ψ e ( r ) ] * E ( r ; ω ) [ ψ j h ( r ) ] * .
n e t | rel = 𝒲 ( in ) ( 1 n e ) 𝒲 ( out ) n e ,
𝒲 ( in ) = 2 π h ¯ | V , ep | 2 n e { 𝒩 ph ( Ω 0 ) [ h ¯ Γ ph / π ( ε ¯ e ε ¯ e h ¯ Ω 0 ) 2 + h ¯ 2 Γ ph 2 ] + [ 𝒩 ph ( Ω 0 ) + 1 ] [ h ¯ Γ ph / π ( ε ¯ e ε ¯ e + h ¯ Ω 0 ) 2 + h ¯ 2 Γ ph 2 ] } + 2 π h ¯ j , j | V , j ; j , eh | 2 ( 1 n j h ) n j h n e [ h ¯ γ eh / π ( ε ¯ e + ε ¯ j h ε ¯ e ε ¯ j h ) 2 + h ¯ 2 γ eh 2 ] ,
𝒲 ( out ) = 2 π h ¯ | V , ep | 2 ( 1 n e ) { 𝒩 ph ( Ω 0 ) [ h ¯ Γ ph / π ( ε ¯ e ε ¯ e h ¯ Ω 0 ) 2 + h ¯ 2 Γ ph 2 ] + [ 𝒩 ph ( Ω 0 ) + 1 ] [ h ¯ Γ ph / π ( ε ¯ e ε ¯ e + h ¯ Ω 0 ) 2 + h ¯ 2 Γ ph 2 ] } + 2 π h ¯ j , j | V , j ; j , eh | 2 ( 1 n e ) ( 1 n j h ) n j h [ h ¯ γ eh / π ( ε ¯ e + ε ¯ j h ε ¯ e ε ¯ j h ) 2 + h ¯ 2 γ eh 2 ] .
n j h t | rel = 𝒲 ¯ j ( in ) ( 1 n j h ) 𝒲 ¯ j ( out ) n j h ,
𝒲 ¯ j ( in ) = 2 π h ¯ j | V j , j hp | 2 n j h { 𝒩 ph ( Ω 0 ) [ h ¯ Γ ph / π ( ε ¯ j h ε ¯ j h h ¯ Ω 0 ) 2 + h ¯ 2 Γ ph 2 ] + [ 𝒩 ph ( Ω 0 ) + 1 ] [ h ¯ Γ ph / π ( ε ¯ j h ε ¯ j h + h ¯ Ω 0 ) 2 + h ¯ 2 Γ ph 2 ] } + 2 π h ¯ , j | V , j ; j , eh | 2 ( 1 n e ) n j h n e [ h ¯ γ eh / π ( ε ¯ e + ε ¯ j h ε ¯ e ε ¯ j h ) 2 + h ¯ 2 γ eh 2 ] ,
𝒲 ¯ j ( out ) = 2 π h ¯ j | V j , j hp | 2 ( 1 n j h ) { 𝒩 ph ( Ω 0 ) [ h ¯ Γ ph / π ( ε ¯ j h ε ¯ j h h ¯ Ω 0 ) 2 + h ¯ 2 Γ ph 2 ] + [ 𝒩 ph ( Ω 0 ) + 1 ] [ h ¯ Γ ph / π ( ε ¯ j h ε ¯ j h + h ¯ Ω 0 ) 2 + h ¯ 2 Γ ph 2 ] } + 2 π h ¯ , j | V , j ; j , eh | 2 ( 1 n e ) ( 1 n j h ) n e [ h ¯ γ eh / π ( ε ¯ e + ε ¯ j h ε ¯ e ε ¯ j h ) 2 + h ¯ 2 γ eh 2 ] ,
| V , ep | 2 = e 2 h ¯ Ω 0 8 π 2 ε 0 ( 1 ε 1 ε s ) d 2 q | , e ( q ) | 2 ( 1 q + q s ) ,
| V , hp | 2 = e 2 h ¯ Ω 0 8 π 2 ε 0 ( 1 ε 1 ε s ) d 2 q | j , j h ( q ) | 2 ( 1 q + q s ) ,
sp ( t ) = | d c , v ( t ) | 2 ε 0 ε b | d 3 r ψ 1 e ( r ) ψ 1 h ( r ) | 2 0 d ω θ [ h ¯ ω c ( t ) ε ¯ 1 e ( ω | t ) ε ¯ 1 h ( ω | t ) ] × h ¯ ω ρ 0 ( ω ) { h ¯ γ eh [ h ¯ ω c ( t ) ε ¯ 1 e ( ω | t ) ε ¯ 1 h ( ω | t ) ] 2 + h ¯ 2 γ eh 2 } ,
| d c , v ( t ) | 2 = e 2 h ¯ 2 2 m 0 G ( T ) [ 1 + c ( t ) G ( T ) ] ( m 0 m e * 1 ) ,
c ( t ) = 1 n 1 e ( t ) ( V 1 , 1 ; 1 , 1 ee V 1 , 1 ; 1 , 1 ee ) + j 1 n j 1 h ( t ) ( V 1 , j 1 ; j 1 , 1 hh V 1 , j 1 ; 1 , j 1 hh ) 1 n 1 e ( t ) V 1 , 1 ; 1 , 1 eh j 1 n j 1 h ( t ) V 1 , j 1 ; j 1 , 1 eh [ 1 n 1 e ( t ) n 1 h ( t ) ] V 1 , 1 ; 1 , 1 eh .
𝒫 loc ( r ; ω ) = 2 | ξ ( r ) | 2 d c , v { d 3 r ψ 1 e ( r ) ψ 1 h ( r ) } × 1 h ¯ lim t { 1 n 1 e ( t ) n 1 h ( t ) ω + i γ 0 Ω ¯ 1 , 1 eh ( ω | t ) } 1 , 1 eh ( t ) ,
E μ ( r ; ω ) = E μ ( 0 ) ( r ; ω ) ω 2 ε 0 c 2 ν 𝒢 μ ν ( r , r 0 ; ω ) 𝒫 ν loc ( ω ) ,
𝒫 loc ( ω ) = 2 d c , v { d 3 r ψ 1 e ( r ) ψ 1 h ( r ) } × 1 h ¯ lim t { 1 n 1 e ( t ) n 1 h ( t ) ω + i γ 0 Ω ¯ 1 , 1 eh ( ω | t ) } 1 , 1 eh ( t ) ,
1 , 1 eh ( t ) = θ ( t ) [ E ( r 0 ; ω ) d c , v ] { d 3 r ψ 1 e ( r ) ψ 1 h ( r ) } * .
E μ ( r ; ω ) = E μ ( 0 ) ( r ; ω ) + 2 ω 2 ε 0 c 2 [ E ( r 0 ; ω ) d c , v ] d c , v | d 3 r ψ 1 e ( r ) ψ 1 h ( r ) | 2 × 1 h ¯ lim t { 1 n 1 e ( t ) n 1 h ( t ) ω + i γ 0 Ω ¯ 1 , 1 eh ( ω | t ) } ν 𝒢 μ ν ( r , r 0 ; ω ) e ^ d ν ,
n e ( t ) f 0 ( ε ¯ e ) 1 exp [ ( ε ¯ e μ e ) / k B T ] + 1 ,
n j h ( t ) f 0 ( ε ¯ j h ) 1 exp [ ( ε ¯ j h μ h ) / k B T ] + 1 ,
ν 𝒜 μ ν ( r 0 ; ω ) E ν ( r 0 ; ω ) = E μ ( 0 ) ( r 0 ; ω )
𝒜 μ ν ( r 0 ; ω ) = δ μ ν 2 ω 2 ε 0 c 2 h ¯ [ 1 f 0 ( ε ¯ 1 e ) f 0 ( ε ¯ 1 h ) ω + i γ 0 Ω ¯ 1 , 1 eh ( ω ) ] | d 3 r ψ 1 e ( r ) ψ 1 h ( r ) | 2 × d c , v 2 [ e ^ d ν ν 1 𝒢 μ ν 1 ( r 0 , r 0 ; ω ) e ^ d ν 1 ] ,
𝒢 μ ν ( r 0 , r 0 ; ω ) = d 2 k ( 2 π ) 2 g μ ν ( k , ω | z 0 , z 0 ) .
E μ ( r ; ω ) = E μ ( 0 ) ( r ; ω ) + 2 ω 2 ε 0 c 2 h ¯ [ ν , ν e ^ d ν 𝒜 ν ν 1 ( r 0 ; ω ) E ν ( 0 ) ( r 0 ; ω ) ] × | d 3 r ψ 1 e ( r ) ψ 1 h ( r ) | 2 d c , v 2 [ 1 f 0 ( ε ¯ 1 e ) f 0 ( ε ¯ 1 h ) ω + i γ 0 Ω ¯ 1 , 1 eh ( ω ) ] ν 1 𝒢 μ ν 1 ( r , r 0 ; ω ) e ^ d ν 1 .
E μ ( k , ω | x 3 ) = E μ ( 0 ) ( k , ω | x 3 ) ω 2 ε 0 c 2 ν g μ ν ( k , ω | x 3 , z 0 ) 𝒫 ν loc ( ω ) .
ν { δ μ ν 2 ω 2 ε 0 c 2 h ¯ [ 1 f 0 ( ε ¯ 1 e ) f 0 ( ε ¯ 1 h ) ω + i γ 0 Ω ¯ 1 , 1 eh ( ω ) ] | d 3 r ψ 1 e ( r ) ψ 1 h ( r ) | 2 d c , v 2 × [ e ^ d ν ν 1 g μ ν 1 ( k , ω | z 0 , z 0 ) e ^ d ν 1 ] } E ν ( k , ω | z 0 ) = E μ ( 0 ) ( k , ω | z 0 ) .
E ( 0 ) ( r ; ω sp ) = E sp e i k 0 ( ω sp ) D 0 c ω sp [ i k ^ 0 β 3 ( k 0 , ω sp ) x ^ 3 k 0 ( ω sp ) ] e i k 0 ( ω sp ) x e β 3 ( k 0 , ω sp ) x 3 ,
k 0 ( ω sp ) = ω sp c ε d ε M ( ω sp ) ε d + ε M ( ω sp ) ,
β 3 ( k 0 , ω sp ) = k 0 2 ( ω sp ) ω sp 2 c 2 ,
E ( 0 ) ( k , ω sp | z 0 ) = δ ( k k 0 ) E sp e i k 0 ( ω sp ) D 0 ( 2 π ) 2 c ω sp × [ i k ^ 0 β 3 ( k 0 , ω sp ) x ^ 3 k 0 ( ω sp ) ] e β 3 ( k 0 , ω sp ) z 0 .
β 0 ( ω ; t ) = ω ε b n spp ( ω ; t ) c [ 1 exp ( h ¯ ω / k B T ) 1 + 1 ] Im [ α spp ( ω ; t ) ] ,
Im [ α spp ( ω ; t ) ] = θ ( t ) ( 2 ε 0 ε b 𝒱 | E ( 0 ) ( r 0 ; ω ) | 2 ) | E ( r 0 ; ω ) d c , v | 2 | d 3 r ψ 1 e ( r ) ψ 1 h ( r ) | 2 × [ 1 n 1 e ( t ) n 1 h ( t ) ] { [ A 2 ( ω ; t ) B 2 ( t ) ] 2 + 4 h ¯ 2 γ 0 2 A 2 ( ω ; t ) [ A 2 ( ω ; t ) + B 2 ( t ) ] 2 + 4 h ¯ 2 γ 0 2 A 2 ( ω ; t ) } [ h ¯ γ 0 Δ 2 ( ω ; t ) + h ¯ 2 γ 0 2 ] ,
Re [ α spp ( ω ; t ) ] = θ ( t ) ( 2 ε 0 ε b 𝒱 | E ( 0 ) ( r 0 ; ω ) | 2 ) | E ( r 0 ; ω ) d c , v | 2 | d 3 r ψ 1 e ( r ) ψ 1 h ( r ) | 2 × [ 1 n 1 e ( t ) n 1 h ( t ) ] { A 4 ( ω ; t ) B 4 ( t ) ] 2 [ A 2 ( ω ; t ) + B 2 ( t ) ] 2 } [ Δ ( ω ; t ) Δ 2 ( ω ; t ) + h ¯ 2 γ 0 2 ] ,
n spp ( ω ; t ) = 1 2 ( 1 + Re [ α spp ( ω ; t ) ] + { 1 + Re [ α spp ( ω ; t ) ] } 2 + { Im [ α spp ( ω ; t ) ] } 2 ) 1 / 2 .
Δ ( ω ; t ) = [ G ( T ) + ε 1 e + ε 1 h h ¯ ω ] 2 + 4 | 1 , 1 eh ( t ) | 2 ,
A 2 ( ω ; t ) = [ h ¯ ω G ( T ) ε 1 e ε 1 h + Δ ( ω ; t ) ] 2 , B 2 ( t ) = 4 | 1 , 1 eh ( t ) | 2 .
α pf ( ω p ; t ) = θ ( t τ ) ( 2 ε 0 ε b 𝒱 | E p | 2 h ¯ ) | E p d c , v | 2 | d 3 r ψ 1 e ( r ) ψ 1 h ( r ) | 2 [ 1 n 1 e ( t ) n 1 h ( t ) ] × { A 2 ( ω ; t ) B 2 ( t ) [ A 2 ( ω ; t ) + B 2 ( t ) ] 2 } { A 2 ( ω ; t ) ω p + i γ eh Ω ¯ 1 , 1 eh ( ω | t ) B 2 ( t ) ω p + i γ eh Ω ¯ 1 , 1 eh ( ω + | t ) } ,
n pf ( ω p ; t ) = 1 2 ( 1 + Re [ α pf ( ω p ; t ) ] + { 1 + Re [ α pf ( ω p ; t ) ] } 2 + { Im [ α pf ( ω p ; t ) ] } 2 ) 1 / 2 .
h ¯ Ω ¯ 1 , 1 eh ( ω ± | t ) = h ¯ ω ± ( t ) [ 1 n 1 e ( t ) n 1 h ( t ) ] V 1 , 1 ; 1 , 1 eh + 1 n 1 e ( t ) ( V 1 , 1 ; 1 , 1 ee V 1 , 1 ; 1 , 1 ee ) + j 1 n j 1 h ( t ) ( V 1 , j 1 ; j 1 , 1 hh V 1 , j 1 ; 1 , j 1 hh ) 1 n 1 e ( t ) V 1 , 1 ; 1 , 1 eh j 1 n j 1 h ( t ) V 1 , j 1 ; j 1 , 1 eh ,
h ¯ ω ± ( t ) = h ¯ ω ± Δ ( ω ; t ) .
𝒫 em ( ω ; t ) | d c , v | 2 ε 0 ε b 0 n 1 e ( t ) n 1 h ( t ) h ¯ γ eh { 1 [ A 2 ( ω ; t ) + B 2 ( t ) ] 2 } | d 3 r ψ 1 e ( r ) ψ 1 h ( r ) | 2 h ¯ ω ρ 0 ( ω ) × { A 2 ( ω ; t ) B 2 ( t ) [ h ¯ ω c ( t ) h ¯ ω ( t ) ] 2 + h ¯ 2 γ eh 2 + A 2 ( ω ; t ) B 2 ( t ) [ h ¯ ω c ( t ) h ¯ ω + ( t ) ] 2 + h ¯ 2 γ eh 2 + A 4 ( ω ; t ) + B 4 ( t ) [ h ¯ ω c ( t ) h ¯ ω ] 2 + h ¯ 2 γ eh 2 } .
V ( r ) = { 0 , 0 x i L i for i = 1 , 2 , 3 , otherwise ,
h ¯ 2 2 m * [ 2 x 1 2 + 2 x 2 2 + 2 x 3 2 + V ( r ) ] ψ ( r ) = ε ψ ( r ) ,
ψ n 1 , n 2 , n 3 ( r ) = 2 L 1 sin [ ( n 1 π L 1 ) x 1 ] 2 L 2 sin [ ( n 2 π L 2 ) x 2 ] 2 L 3 sin [ ( n 3 π L 3 ) x 3 ] ,
ε n 1 , n 2 , n 3 = h ¯ 2 2 m * [ ( n 1 π L 1 ) 2 + ( n 2 π L 2 ) 2 + ( n 3 π L 3 ) 2 ] ,
λ α e ( ω | t ) = λ α h ( ω | t ) = { 1 2 ( h ¯ ω + [ G ( T ) + ε α e + ε α h h ¯ ω ] 2 + 4 | α , α eh ( t ) | 2 ) if h ¯ ω G ( T ) + ε α e + ε α h 1 2 ( h ¯ ω [ G ( T ) + ε α e + ε α h h ¯ ω ] 2 + 4 | α , α eh ( t ) | 2 ) if h ¯ ω G ( T ) + ε α e + ε α h ,
n 1 , n 2 , n 3 ; n 1 , n 2 , n 3 e ( q ) = n 1 , n 2 , n 3 ; n 1 , n 2 , n 3 h ( q ) = 𝒬 n 1 , n 1 ( 1 ) ( q 1 ) 𝒬 n 2 , n 2 ( 2 ) ( q 2 ) 𝒬 n 3 , n 3 ( 3 ) ( q 3 ) ,
𝒬 n j , n j j ( q j ) = ( 2 L j ) 0 L j d x j e i q j x j sin [ ( n j π L j ) x j ] sin [ ( n j π L j ) x j ] .
d 3 r ψ n 1 , n 2 , n 3 e ( r ) ψ n 1 , n 2 , n 3 h ( r ) = δ n 1 , n 1 δ n 2 , n 2 δ n 3 , n 3 .
d c , v = e 2 h ¯ 2 2 m 0 G ( T ) ( m 0 m e * 1 ) .
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.