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

Phonon-assisted photoluminescence from a semiconductor quantum dot with resonant electron and phonon subsystems

Open Access Open Access

Abstract

We present a theory of phonon-assisted photoluminescence from a semiconductor quantum dot (QD) whose electron and phonon subsystems are resonantly coupled via the polar electron–phonon interaction. We show that the resonance-induced renormalization of the QD energy spectrum, leading to the formation of the polaron-like states, can be performed exactly in terms of the arbitrarily degenerate states of electron–hole pairs and the phonon modes of equal energies. Using the model of QDs with finite potential barriers for electron and holes leads to new selection rules of interband optical transitions and the three-particle interaction describing simultaneous absorption and/or emission of a photon and a phonon. We also derive a simple expression for the differential cross section of the stationary, low-temperature photoluminescence, which allows the fundamental parameters of the polaron-like excitations to be readily extracted from the frequency-resolved experimental spectra. In particular, the energies of the excitations and the coherence relaxation rates of the optical transitions resulting in their generation and recombination are shown to be directly given by the positions and widths of the photoluminescence peaks. The developed theory complements the existing experimental techniques of studying the phonon-assisted photoluminescence from individual nanocrystals.

© 2014 Optical Society of America

1. Introduction

The secondary-emission spectroscopy is the powerful optical tool for studying both individual quantum dots (QDs) and their ensembles [18]. It enables one to determine the chemical composition of the QD material [914], measure the mechanical stress inside a QD [1519], investigate the properties of the QD surface [2023], and study various aspects of the electron–phonon interaction inside QDs [2426]. Many of the unique physical properties [27, 28] of semiconductor QDs underlying their applications in photonics [29, 30] were determined using the experimental methods of the secondary-emission spectroscopy.

Among the most important interactions in semiconductor QDs is the electron–phonon coupling, which enhances the dephasing rates of spectroscopic transitions [3134] and reduces the lifetimes of electronic excitations [3539]. This interaction also leads to the formation of hybrid polaron-like states inside QDs with resonant electron and phonon subsystems [4043], causing a significant modification of the QDs’ energy spectra. The vibrational resonance required for the modification to occur is realized, for instance, where the energy of a phonon mode (supported either by the QD or the host material [44, 45]) coincides with the energy spacing between a pair of the quantum-confined electron states of the QD. Often, these are the longitudinal optical (LO) phonons and the polar electron–phonon interaction that couple the states and induce the resonance [43,4651]. The energy splitting of the resulting polaron-like excitations is typically in the millielectronvolt range and can therefore be experimentally resolved only at cryogenic temperatures.

The formation of the polaron-like excitations in the above circumstances have been theoretically described for a variety of semiconductors by several research groups [43, 5058]. In particular, Hameau et al. [51] proposed a model of everlasting resonant polarons in InAs QDs to explain the experimental data on the QD intraband transitions obtained with the far-infrared magnetospectroscopy [46, 47, 49, 51]. The multiphonon Raman scattering and polaron effects in semiconductor nanocrystals were taken into account by Pokatilov et al. [53] and Miranda et al. [54, 57] to describe the Raman spectrum of spherical CdSe QDs. Itoh et al. [43] developed a theory of polaron and exciton-phonon complexes to explain the shape of the exciton luminescence spectra measured under the resonant size-selective excitation of CuCl nanocrystals. Another model of the polaron-like excitations, suggested by Fedorov et al. [50], was employed to interpret the renormalization of the lowest-energy exciton states in CuCl QDs embedded in the NaCl matrix, which were studied under the resonant two-photon excitation using the secondary-emission spectroscopy. This model was recently used to describe the transformation of the QD energy spectrum due to the resonant polar electron–phonon coupling and to study how this transformation is manifested in the stationary photoluminescence spectra [40].

This paper presents a theory of phonon-assisted secondary emission from a single QD in the regime of vibrational resonance. In contrast to previous works, we are considering a vibrational resonance in a QD with finite potential barriers for the confined carriers. Owing to the barriers finiteness, the selection rules of interband optical transitions are modified, and transitions to some electron–hole pair states that cannot be optically excited in the QDs with impenetrable boundaries become dipole allowed. Another important consequence of the QD model adopted is the emergence of a specific three-particle interaction between the optical radiation and the electron–phonon subsystem of the QD. This interaction leads to a multi-peak structure of the photoluminescence and excitation photoluminescence spectra, whose analysis allows one to estimate the strength of the electron–phonon interaction.

The paper is organized as follows. We begin in Section II by extending the Hamiltonian formalism employed earlier for the treatment of the vibrational resonance in the QD being an infinitely high potential well for the confined charge carriers [40,52] to the case of a spherical QD with a penetrable surface. This allows us to consider the three-particle interaction—involving electrons–hole pair, phonons, and photons—essential for the study. Important here is that this interaction is readily observable in the secondary emission spectra but absent in the model of infinitely high potential barriers. Section III presents a stationary solution to the generalized master equation for the reduced density matrix, which describes the process of photoluminescence mediated by the decay of the polaron-like excitations of the QD. The energies and wave functions of the excitations are calculated analytically by diagonalizing the Hamiltonian of the system with respect to the electron–phonon interaction in the Appendix. Section IV summarizes our results and concludes the paper.

2. Theory of vibrational resonance

Our approach to the theoretical treatment of vibrational resonance in a semiconductor QD is based on the use of the Hamiltonian formalism that was initially developed by Levinson and Rashba [59] for bulk materials and later extended by Fedorov et al. [40, 50, 52] for low-dimensional structures. Without loss of generality, we restrict our consideration to the polar electron–phonon interaction between the charge carriers and longitudinal optical (LO) phonons confined inside a semiconductor QD with a high degree of ionicity. This interaction is relatively easy to study experimentally due to its greater strength as compared to the strengths of the other kinds of electron–phonon coupling [24, 26, 41, 43, 49, 51]. The following theory, with the exception of Eqs. (2)(9), is applicable to any kind of electron–phonon interaction leading to the formation of polaron-like states in quantum nanostructures.

2.1. Hamiltonian formalism

The phonon-assisted photoluminescence from a semiconductor QD can occur according to four different physical scenarios, in which phonons are either absorbed or emitted upon the optical excitation or radiative relaxation of the QD’s electronic subsystem [40]. We describe the electron (e) and phonon (ph) subsystems of the QD, their excitation with light (L), the radiation field (r), and the interactions between the electromagnetic fields and the QD using six Hamiltonians: (i) the free electron–hole pair Hamiltonian He expressed through energies, Ep, creation operators, ap+, and annihilation operators, ap, of the pairs; (ii) the Hamiltonian Hph of the dispersionless LO phonon modes of energy Ω described by operators bq+ and bq; (iii) the radiation-field Hamiltonian Hr, which is expressed through the energies h̄ωλ of the electromagnetic field modes λ and the modes’ creation, cλ+, and annihilation, cλ, operators; (iv) the polar electron–phonon coupling Hamiltonian He,ph; (v) the Hamiltonian He,L of the QD interaction with the classical excitation field; and (vi) the Hamiltonian He,r of the interaction between the QD and the quantum radiation field. The energy of an electron–hole pair in state |p〉 is comprised of three terms, Eg + Epe + Eph, which are the band gap of bulk semiconductor and the quantum-confinement energies of electrons and holes in the states characterized by the sets of quantum numbers pe and ph.

Let us focus on the QD in the strong confinement regime and describe its electron subsystem using a simple two-band model of semiconductor material [2426]. Using the parabolic model of the valence band and neglecting the Coulomb interaction allow us to derive simple formulas for rapid plotting of photoluminescence spectra and checking the simulation results. Since EgΩ for most semiconductors with a high degree of ionicity, the electron–phonon coupling induces only intraband transitions of the confined charge carriers, whereas the interaction of the QD’s electronic subsystem with electromagnetic fields leads solely to interband transitions. The Hamiltonian operators describing these interactions are of the forms [40]

He,ph=p1,p2qVp2;p1(q)ap2+ap1bq+H.c.,
He,L=pVp;0(L)ap++H.c.
and
He,r=p,λVp;0(λ)ap+cλ+H.c.,
where Vp2;p1(q), Vp;0(L) and Vp;0(λ) are the matrix elements.

The strength of the polar electron–phonon interaction is determined by the matrix element [24] Vp2;p1(q)=e(φph1,ph2(q)δpe2,pe1φpe2,pe1(q)δph2,ph1), where −e is the charge of the electron, φph1,ph2(q) and φpe2,pe1(q) are the matrix elements of the electric field induced by the LO phonons of mode q; and δph2,ph1 and δpe2,pe1 are the products of Kronecker deltas for the respective pairs of quantum numbers from sets ph1 and ph2 for holes and pe1 and pe2 for electrons.

The direct creation of an electron–hole pair by the classical excitation field and the subsequent annihilation of the pair due to its interaction with the vacuum radiation field are described with the matrix elements Vp;0(L)=eELDcvFpϕ(t)exp(iωLt) and Vp;0(λ)=ie2πh¯ωλ/(εV)DcvFp, where ϕ(t) is the complex-valued envelope of the excitation pulse of the peak field strength EL, Dcv is the dipole matrix element calculated on the Bloch wave functions at the Brillouin zone center, Fp is the overlap integral calculated on the envelope wave functions of electrons and holes, ε is the high-frequency permittivity of the bulk semiconductor, and V is the normalization volume.

2.2. A spherical QD with a penetrable surface

We shall eventually be interested in calculating the intensity of the QD secondary emission. This will require explicit expressions for the matrix elements entering Eq. (1). To find the forms of these elements, we now specify the model of the QD.

We first assume that the QD semiconductor has Td or Oh symmetry. This enables us to relate the dipole matrix element to Kane’s parameter P of the bulk material through the bandgap energy via |Dcv|=2P/Eg [40]. We also consider our QD to be a sphere of radius R embedded in a dielectric host. Then the states of electrons, holes and phonons are characterized by sets {n, l, m} of three quantum numbers: the principal quantum number n, the angular momentum l, and its projection m. The overlap integral in this case is of the form [60]

Fp=2δle,lhδme,mh0Rne,leRnh,lhx2dx,
where Rnl(x) is the radial part of the envelope wave function.

Using the continual model of the dispersionless LO phonons, we next analytically calculate the electric field potential induced by the phonons inside and outside the QD [40]. The matrix elements of this field calculated on the envelope wave functions of electrons and holes are given by

φpk2,pk1(q)=4(2lq+1)(2lk1+1)h¯Ω(2lk2+1)ε*RClq0,lk10lk20Clqmq,lk1mk1lk2mk2nk2lk2,nk1lk1nqlq,
where q = {nq, lq, mq} is the set of phonon quantum numbers, pki = {nki, lki, mki} (i = 1, 2) is the set of quantum numbers characterizing the state of electrons (k = e) or holes (k = h), ε* = ε0ε/(εε0), ε0 is the low-frequency permittivity, Clqmq,lk1mk1lk2mk2 is the Clebsch–Gordan coefficient,
nk2lk2,nk1lk1nqlq=0nqlqRnk1lk1Rnk2lk2x2dx,
and nqlq (x) is the radial part of the electric-field potential induced by the phonon mode.

The forms of the radial functions Rnl(x) and nqlq (x) are determined by the profiles and heights of the potential barriers that need to be overcome by electrons, holes and phonons to leave the QD. These barriers are usually finite and unequal for electrons and holes confined by a semiconductor QD in a dielectric host. By describing them with the step functions of heights Ve and Vh, we obtain the following functional dependency [60]:

Rnklk=1Anklk×{κlk(αnklks)jlk(αnklkdx),x1jlk(αnklkd)κlk(αnklksx),x>1,
in which
Anklk=κlk+1(αnklks)κlk1(αnklks)jlk2(αnklkd)jlk+1(αnklkd)jlk1(αnklkd)κlk2(αnklks),
αnklkd=(R/h¯)2mkdEnklk,
αnklks=(R/h¯)2mks(VkEnklk);
jl(x) and κl(x) are, respectively, the spherical and modified spherical Bessel functions; and mkd and mks are the effective masses of electrons (k = e) and holes (k = h) inside the QD and in the surrounding medium. The energy spectra, Enklk, of electrons and holes are determined by the secular equation
mksαnklkdjlk(αnklkd)jlk(αnklkd)=mkdαnklksκlk(αnklks)κlk(αnklks),
with primes denoting the first derivatives of the functions.

Since our QD is embedded in the dielectric that does not support optical phonon modes, its surface may be considered impenetrable for the confined LO phonons. The radial dependency of the phonon potential in this case is given by [6163]

nqlq=1ξnqlqjlq+1(ξnqlq)×{jlq(ξnqlqx),x10,x>1,
where ξnl is the nth root of equation jl(x) = 0.

The overlap integral and the radial part of the matrix element φpk2,pk1(q) can now be calculated by substituting Eqs. (5) and (8) into Eqs. (2) and (4) to obtain

FpFleme,lhmhnenh=2δle,lhδme,mhAneleAnhlh(κle(αneles)κlh(αnhlhs)01jle(αneledx)jlh(αnhlhdx)x2dx+jle(αneled)jlh(αnhlhd)1κle(αnelesx)κlh(αnhlhsx)x2dx)
and
nk2lk2,nk1lk1nqlq=κlk1(αnk1lk1s)κlk2(αnk2lk2s)ξnqlqjlq+1(ξnqlq)Ank1lk1Ank2lk2×01jlq(ξnqlqx)jlk1(αnk1lk1dx)jlk2(αnk2lk2dx)x2dx.

Equation (9a) shows that if a QD provides finite potential barriers for the confined carriers then the overlap integrals of the product of the envelope wave functions of electrons and holes with different principal quantum numbers (nenh) are nonzero. In the case of infinite barriers this integral is given by the triple product δne,nhδle,lhδme,mh, vanishing for nenh. Hence, the finiteness of the potential barriers drastically modifies the selection rules of interband optical transitions, allowing some transitions that are dipole-forbidden in QDs with impenetrable boundaries.

2.3. Formation of polaron-like states

The vibrational resonance occurs where the energy of one or several LO phonon modes is close to the energy gap between a pair of the electron–hole states, i.e. where Ω ≈ Ep2Ep1. The energy of the QD with one phonon and an electron in the lower state is equal in this situation to the energy of the QD without phonons and an electron in the upper resonant state. The degeneracy of the QD states is removed by the polar electron–phonon interaction, which leads to their splitting into several hybrid polaron-like states of different energies [40, 50, 52]. The number and degeneracy of the resulting states depend on the degeneracy orders of the electron, hole and phonon states involved in the vibrational resonance. The splitting is schematically illustrated by Fig. 1.

 figure: Fig. 1

Fig. 1 Formation of polaron-like states from a pair of nondegenerate electron–hole states coupled via one, two, three or four degenerate LO phonon modes of energy Ω. The electron–phonon interaction transforms the degenerate state of the QD into two, three, four or five nondegenerate polaron-like states.

Download Full Size | PDF

The calculation of the energies and wave functions of the polaron-like states requires diagonalization of the Hamiltonians in Eq. (1) with respect to the electron–phonon interaction. This diagonalization is carried out in two steps. With the first step, we exclude the diagonal part of the electron–phonon coupling using the transformation [40]

H˜=U+HU,
where
U=exp(p,q(Φp;p(q)bqH.c.)ap+ap)
and Φp;p(q)=Vp;p(q)/(h¯Ω). If we keep only the linear terms in bq and bq+, then the energies of the transformed states become
E˜p=Epqh¯Ω|Φp;p(q)|2
and the new Hamiltonians take the forms [52]
H˜e,ph=p1p2qVp2;p1(q)ap2+ap1bq+H.c.,
H˜e,L=pVp;0(L)(1+q(Φp;p(q)bqH.c.))ap++H.c.,
and
H˜e,r=p,λVp;0(λ)(1+q(Φp;p(q)bqH.c.))ap+cλ+H.c.

The transformed Hamiltonians responsible for the coupling of the polaron-like states to the excitation light [Eq. (11c)] and radiation field [Eq. (11d)] are seen to contain additional terms describing three-particle interactions. These interactions lead to the creation or annihilation of a polaron-like quasiparticle with the simultaneous absorption and/or emission of one photon and one phonon, as illustrated by the Feynman diagrams in Fig. 2. The four processes represented by the diagrams mediate the phonon-assisted secondary emission, which can be observed in the QD photoluminescence spectra.

 figure: Fig. 2

Fig. 2 Feynman diagrams describing [(a) and (b)] creation and [(c) and (d)] annihilation of a polaron-like quasiparticle of energy p with the simultaneous (a) absorption of a photon, h̄ωL, and a phonon, Ω, (b) absorption of a photon and emission of a phonon, (c) emission of a photon, h̄ωλ, and absorption of a phonon, and (d) emission of a photon and a phonon.

Download Full Size | PDF

In Fig. 3 we plot the absolute values of matrix elements Φp,p and overlap integrals Fp as functions of potential barrier heights for electrons and holes in a spherical InAs QD with R = 9.14 nm (the barrier heights are assumed to be related as Ve = 1.35Vh). The plots correspond to the vibrational resonance between the orbitally nondegenerate states |p1〉 = |1, 0, 0;1, 0, 0〉 and |p2〉 = |1, 0, 0; 2, 0, 0〉 of the valence band and a phonon mode of quantum numbers q = {1, 0, 0}. The Hamiltonian terms of the form Fp(1+Φp;p(q)bqΦp;p(q)*bq+)ap+, which describe the interaction of the polaron-like states with excitation light, are seen to be all nonzero for finite potential barriers. Since Φp1,p1 and Fp2 (shown by dashed curves) vanish when the barriers are infinitely high, the phonon-assisted photoluminescence in this case is absent.

 figure: Fig. 3

Fig. 3 Matrix elements and overlap integrals as functions of potential barrier heights for electrons and holes in a spherical InAs QD with R = 9.14 nm. The material parameters are: ε0 = 15.15, ε = 12.25, Eg = 418 meV, Ω = 29.5 meV, me = 0.0219m0, and mh = 0.43m0 [64] (m0 is the free-electron mass). For other parameter refer to the text.

Download Full Size | PDF

With the second step, we exclude from Eqs. (11b)(11d) the nondiagonal part of the electron–phonon coupling corresponding to the states in the vibrational resonance using the transformation Ĥe,ph = S+e,phS. In order to construct the unitary operator S performing the transformation, we solve the standard eigenvalue problem Ĥe,ph|ψ̂〉 = |ψ̂〉, where eigenstates |ψ̂〉 are related to the eigenstates |ψ̃〉 of the Hamiltonian e,ph through the Hermitian conjugate of S via |ψ̂〉 = S+|ψ̃〉. For the general case in which the resonance condition is simultaneously satisfied for k phonon modes and the polaron-like states of degeneracies k1 and k2 (subscripts 1 and 2 designate the lower and upper states, respectively), the eigenvalue problem leads to the algebraic equation of order s = kk1 + k2. This equation can be solved exactly only for s ≤ 4, while also admitting analytical solutions in many situations of practical interest for s > 4.

The spherical symmetry of the QD enables analytical treatment of the formation of the polaron-like states in the quite general case of arbitrary degeneracies of the electron and hole states with respect to the projections of the angular momentum. In the Appendix, we use the selection rules of the electron–phonon coupling to construct the Hamiltonian H˜e,ph(k), carry out its diagonalization, and obtain the energies and wave functions of the polaron-like excitations [see Eq. (18)(23)] for a pair of arbitrarily degenerate hole states resonant to an arbitrary number of dispersionless phonon modes. We also show that the diagonalization of the electron–phonon Hamiltonian in the case of degenerate electron–hole-pair states can be reduced to a number of identical diagonalization problems for the pairs of nondegenerate states coupled via eligible phonon modes (with nonzero matrix elements). In Section III, we consider an example leading to such a basic problem, in which the vibrational resonance occurs between a pair of orbitally nondegenerate hole states and two phonons of zero angular momentum.

2.4. Polaron–photon interaction

The formation of the polaron-like states modifies interactions of the QD with excitation light and emitted photons. This modification is described by the same transformation of the Hamiltonians e,L and e,r that was used earlier to eliminate the nondiagonal part of the electron–phonon coupling. The results of the Appendix show that the matrix elements of the transformed Hamiltonian describing the optical excitation of the polaron-like states are given by

H^e,L(k)=(000H1(L,k)000H2(L,k)000Hk+1(L,k)H1(L,k)*H2(L,k)*Hk+1(L,k)*0),
where
H1(L,k)=Vp2;0(L)S1;1(k)+Vp1;0(L)ν=1kΦp1;p1(qν)S1+ν;1(k),
H2(L,k)=Vp2;0(L)S1;2(k)+Vp1;0(L)ν=1kΦp1;p1(qν)S1+ν;2(k),
and
Hn(L,k)=Vp1;0(L)ν=1kΦp1;p1(qν)S1+ν;n(k)
for 3 ≤ nk + 1.

The matrix representation of the modified interaction with the emitted photons is

H^e,r(k)=(000H1,q1(r,k)H1,q2(r,k)H1,qk(r,k)000H2,q1(r,k)H2,q2(r,k)H2,qk(r,k)000Hk+1,q1(r,k)Hk+1,q2(r,k)Hk+1,qk(r,k)H1,q1(r,k)*H2,q1(r,k)*Hk+1,q1(r,k)*000H1,q2(r,k)*H2,q1(r,k)*Hk+1,q2(r,k)*000H1,qk(r,k)*H2,qk(r,k)*Hk+1,qk(r,k)*000),
where
H1,q(r,k)=Vp1;0(λ)S1+q;1(k)+Vp2;0(λ)Φp2;p2(q)S1;1(k),
H2,q(r,k)=Vp1;0(λ)S1+q;2(k)+Vp2;0(λ)Φp2;p2(q)S1;2(k),
and
Hn,q(r,k)=Vp1;0(λ)S1+q;n(k)
for 3 ≤ nk + 1.

The rows from top to bottom and columns from left to right in Eqs. (12a) and (13a) correspond to states |1(k)|0λ|0q, |2(k)|0λ|0q,,|k+1(k)|0λ|0q, and |0(k)〉|0λ〉|0q〉 for the interaction with the excitation light, and to states |1(k)|0λ|0q,,|k+1(k)|0λ|0q, |0(k)〉|1λ〉 |1q1〉,..., |0(k)〉|1λ〉|1qk〉 for the interaction with the emitted photons. Here |n(k)|ψ^n are the polaron-like states of energies n(k) given in Eq. (23); |0(k)〉 denotes the vacuum of the polaron-like excitations; and |Nλ〉 and |Nq〉 represent the states with N photons in mode λ and N phonons in mode q.

It should be noted that the use of the infinite potential well model of the QD in our previous work [40] results in the possibility of generation and annihilation of the polaron-like excitations only in states |1(k) and |2(k), whereas the three-particle processes involving phonons enable all polaron-like states to directly interact with light. Indeed, the terms describing the three-particle interaction in Eqs. (11c) and (11d) are proportional to the product FpVp;p(q), where the set of quantum numbers p = {ne, le, me; nh, lh, mh} describes the state of the electron–hole pair. If the QD is modeled by an infinitely deep potential well, then the first factor in this product (equal to δne,nhδle,lhδme,mh) ensures that only electrons and holes of equal quantum numbers are generated or can recombine. The second factor in this case is identically zero, implying that the triple interaction is absent. On the other hand, FpVp;p(q)0 for nenh in the QD with a penetrable surface, in which case Eq. (3) shows that the simultaneous emission or absorption of a photon and a phonon by an electron–hole pair is possible for lq = 0, 2, 4,..., 2l and mq = 0.

It should be noted that the above calculations can be performed in a similar fashion for nonspherical QDs. The lowering of the QD symmetry would reduce the degeneracy degree of electron and hole states, and modify the selection rules of interband optical transitions. With the reduction of the states degeneracy, the treatment of vibrational resonance becomes simpler but the overall calculation scheme remains the same: the renormalization of energy spectrum induced by the electron–phonon interaction is followed by the modification of the Hamiltonian describing the interaction of phonon-polaritons with light.

3. Phonon-assisted secondary emission

We now use the results of the previous section to calculate the differential cross section of the phonon-assisted photoluminescence within the frame of the density matrix formalism [65]. To do so, we assume that the optical excitation of the QD is sufficiently weak for the depletion of the ground electron–hole-pair state and that the temperature of the system is low enough (i.e. kBTΩ) for the spontaneous generation of the optical phonons. The first assumption enables the employment of the perturbation theory whereas the second allows us to neglect the processes of secondary emission accompanied by the emission of the phonons [see the diagrams in Figs. 2(b) and 2(d)].

If the vibrational resonance occurs between k phonon modes and a pair of nondegenerate electron–hole-pair states, then the minimal sufficient basis of our system is formed by the following 2k + 2 eigenstates of Hamiltonian Ĥ:

|i=|0|0q|0λ,
|1=|1(k)|0q|0λ,,|k+1=|k+1(k)|0q|0λ,
|f1=|0|1q1|1λ,,|fk=|0|1qk|1λ.
Here |i〉 is the stationary ground state of the QD; |1〉,...,|k + 1〉 are the excited QD states with zero photons and phonons and the polaron-like quasiparticles of energies 1(k),,k+1(k); and |f1〉,..., |fk〉 are the final states of energy (ωλ + Ω) corresponding to the vacuum of the polaron-like excitations, one emitted photon of frequency λ, and one LO phonon in modes q1,..., qk.

The process of the QD photoluminescence involves several transitions between the states of the basis. The excitation light first transfers the QD from the ground state to a superposition of the excited states by creating one or several polaron-like excitations. As we mentioned earlier, the creation can occur either directly (i.e., purely by light) in states |1(k) and/or |2(k) or in any of the excited states with the emission of a phonon [see Eqs. (22a) and (23c)]. The excited polaron-like states can then nonradiatively convert to one another due to the thermal interaction with a bath at rates which we denote as ζνμ. The polaron-like excitations finally annihilate with the emission of photons h̄ωλ, and the system transits to one of the final states. Unlike the creation, the annihilation can directly produce a phonon which is already present in any of the excited states, as well as occur indirectly in the first two polaron-like states with the simultaneous emission of secondary radiation and creation of a phonon.

By perturbatively solving the generalized master equation [40, 65]

ρμν(t)t=1ih¯[H^(t),ρ(t)]μνγμνρμν(t)+δμνννζννρνν(t),
for the reduced density matrix to the second orders in the interaction energies Vp;0(L), Vp2;p1(q) and Vp;0(λ), we find that the differential cross section of the stationary (ϕ = 1) photoluminescence is given by the expression
d2σdΘdωλ=C(ωλ)η=1k(μ=1k+12γiμγμμ2γ^iμ+γphΔLμ2+γiμ2|dμ(L,k)|2|dμ,qη(λ,k)|2Δλμ2+γfμ2+ν=2k+1μ=1,(μ2)ν12ζνμγiνγμμγνν2γνfΔLμ2+γiμ2|dμ(L,k)|2|dν,qη(λ,k)|2Δλν2+γfν2),
where C(ωλ) = 4(eDcv ωλ)4/(πc42), c is the speed of light in vacuum, ΔLμ=ωμ(k)ωL is the detuning of the excitation light frequency from the frequency ωμ(k)=μ(k)/h¯ of the μth polaron-like state, Δλν=ων(k)ωλΩ, and the modified matrix elements are given by
dn(L,k)=Fp2S1;n(k)+Fp1ν=1kΦp1;p1(qν)S1+ν;n(k),
dn,q(λ,k)=Fp1S1+q;n(k)+Fp2Φp2;p2(q)S1;n(k).
There are two kinds of spectroscopic experiments described by the dependency of the differential cross section in Eq. (16) on excitation and detection frequencies, ωL and ωλ. By fixing ωL and varying ωλ one measures a simple photoluminescence spectrum whereas if ωλ is fixed and ωL is varied one gets the excitation-photoluminescence spectrum. For given frequency detunings Δ and Δλμ, the relative importance of the terms in Eq. (16) is set by: (i) the pure dephasing rates γ̂; (ii) the population decay rate γμμ of the electron–hole levels μ = 1, 2,..., k + 1; (iii) the coherence relaxation rates γ = γμμ/2 + γ̂ and γ = γμμ/2 + γph/2 + γ̂ of the optical transitions resulting in generation and recombination of the polaron-like excitations; (iv) the rates ζνμ of transitions |μ(k)|ν(k); and (v) phonon lifetime γph1. The matrix elements in Eqs. (17a) and (17b) describe, respectively, the direct and phonon-assisted generation (annihilation) of the polaron-like excitations.

As an illustration of the developed theory, we consider the vibrational resonance coupling the orbitally nondegenerate states |p1〉 = |1, 0, 0;1, 0, 0〉 and |p2〉 = |1, 0, 0;2, 0, 0〉 in the valence band. In this case only phonons with zero angular momentum are involved in the resonance, and we restrict our analysis to a pair of them (s = 2) characterized by the quantum numbers q1 = {1, 0, 0} and q2 = {2, 0, 0} (see Fig. 1). We consider InAs QD with Ve = 2.7 eV and Vh = 2 eV, and the rest of material parameters the same as in Fig. 3. We also neglect transitions between the polaron-like states by setting ζνμ = 0 for simplicity. The rest of the relaxation parameters depends on many factors (including the geometry, fabrication quality, and excitation conditions of the sample) and is taken to be γ̂ = 10 μeV, γμμ = 20 μeV, γ = 50 μeV and γph = 60 μeV.

Figure 4 shows five photoluminescence spectra of InAs QD in the spectral range of the one-phonon Stokes sideband. The upper three curves represent the case of resonant excitation at the energies of 530.6, 533.7, and 536.8 meV, whereas the lower two curves correspond to the nonresonant excitation at 527.6 and 539.9 meV. All five photoluminescence spectra are seen to have three peaks each, corresponding to the radiative decay of the polaron-like QD states |1(2), |2(2), and |3(2). It is seen that the blue, orange and red spectra have main peaks at resonant energies h¯(ω2(2)Ω)=501.1meV, h¯(ω3(2)Ω)=504.2meV and h¯(ω1(2)Ω)=507.3meV. The amplitudes of the additional two peaks of the spectra are much smaller than the amplitudes of the main peaks. The nonresonant excitation, represented by the wine-color and green spectra, significantly reduces the intensity of secondary emission while making the amplitudes of the three peaks comparable. The widths of the phonon-assisted photoluminescence peaks are all set by the relaxation parameters γ and equal to 100 μeV.

 figure: Fig. 4

Fig. 4 Photoluminescence spectra of InAs QD with a pair of nondegenerate electronic states coupled through two phonon modes. The excitation frequencies are: h¯ω2(2)=530.6meV, h¯ω3(2)=533.7meV, and h¯ω1(2)=536.8meV; V(2) = 3.1 meV. For material parameters refer to the text.

Download Full Size | PDF

The excitation spectra of photoluminescence are plotted in Fig. 5. The upper spectra correspond to the resonant detection at the energies of 501.1, 504.2 and 507.3 meV, whereas the lower spectra are for the case of nonresonant detection at 498.1 and 510.4 meV. One can see that each of the resonant-detection spectra has one pronounced main peak and two smaller additional peaks. The main peaks are centered at the resonance frequencies ω1(2), ω2(2) and ω3(2) and characterize the decay of the respective polaron-like states, whereas the additional peaks appear due to the photoluminescence mediated by the decay of the states detected nonresonantly.

 figure: Fig. 5

Fig. 5 Excitation photoluminescence spectra of InAs QD with a pair of nondegenerate electronic states coupled through two phonon modes. The detection frequencies are shown near the spectra. The parameter values and resonant frequencies ωj(2) (j = 1, 2, 3) are the same as in Fig. 4.

Download Full Size | PDF

The nonresonant-detection peaks are all of approximately equal intensities, and the widths of all peaks in the figure are 40 μeV. Notice also that had we taken into account the transitions between different polaron-like states (described by factors ζνμ), the intensity of the peak corresponding to the lowest-energy polaron-like state |2(k) would have increased.

The analysis of Eq. (16), facilitated by the above examples, reveals that the spectra of the low-temperature photoluminescence from a semiconductor QD with resonant electron and phonon subsystems is the sum of Lorentzians centered at frequencies ωμ(k) and ωμ(k)Ω. The full widths at half maxima of the Lorentzians are equal to 2γ for the photoluminescence spectra and 2γ for the excitation photoluminescence spectra. By extracting the relaxation rates γ and γ from the experimental spectra, one can calculate the phonon lifetime as [2(γγ)]−1. The nonstationary secondary emission from a QD in the regime of vibrational resonance provides more information on the relaxation parameters and will be theoretically analyzed in our subsequent papers.

Our theory shows that photoluminescence and excitation photoluminescence spectra measured in the regime of vibrational resonance significantly differ from the ordinary spectra of phonon-assisted secondary emission due to the formation of the polaron-like quasiparticles. Indeed, in the absence of vibrational resonance, there would be a single luminescence peak red-shifted from the excitation frequency by the LO-phonon energy. There would also be a similar red-shifted peak in the excitation photoluminescence spectrum corresponding to the optical transition to a certain resonant electron-hole pair state. In sharp contrast to this, in the case of vibrational resonance, both spectra have three peaks each in the spectral range of the one-phonon sideband even at cryogenic temperatures. The energy spacing between the adjacent peaks in the situation of the exact resonance equals V(2) [see Eqs. (23a) and (23b)], i.e. directly related to the strength of the electron–phonon coupling.

4. Conclusions

We have theoretically studied the process of QD photoluminescence involving the emission of the dispersionless optical phonons that are confined by the QD and resonant to a pair of its electron states. To describe this process, we had to abandon the infinite-potential-well model of the QD and assume that its surface is penetrable for the confined charge carriers. This assumption changes the standard selection rules and allows the polaron-like states arising from the resonant electron–phonon coupling to decay with the emission of phonons in spherical QDs. The most important consequence of modelling the QD as finite potential barriers for electrons and holes is the emergence of the three-particle interaction between photons, phonons and electrons (holes). Allowing for this interaction drastically modifies the photoluminescence and excitation photoluminescence spectra, resulting in a fine structure of the one-phonon Stokes sideband. The analysis of this structure allows one to estimate the strength of the electron–phonon interaction inside the QD. Using the density matrix formalism, we then derived an exact expression for the differential cross section of the stationary, low-temperature photoluminescence. Being applicable to the cases of arbitrarily degenerate electron states and an arbitrary number of phonons, our expression describes a wide variety of practical situations. It may prove useful to experimentalists in the analysis of the frequency-resolved photoluminescence spectra and determination of the fundamental parameters of the polaron-like excitations in semiconductor QDs.

Appendix: The S-matrix

Consider the renormalization of the electron–hole-pair energy spectrum of a spherical QD in the regime of vibrational resonance using the S-matrix. For the sake of definiteness, we focus on the resonance in the valence band for a pair of (2l1 + 1)- and (2l2 + 1)-fold degenerate hole states of angular momenta l1 and l2 (subscripts 1 and 2 mark, respectively, the lower and upper states with arbitrary principal quantum numbers), and assume that the state of electrons is nondegenerate. Then the part of Hamiltonian describing the electron–phonon interaction can be represented by a (2l2 + 1)×(2l2 + 1) quasi-diagonal block matrix

(l1,l2)=(H(l1,l2,l2)0000H(l1,l2,l21)0000H(l1,l2,1l2)0000H(l1,l2,l2)),
each block of which describes the coupling of the upper state with a particular projection of l2 to all possible lower states characterized by different projections of l1.

The S-matrix has a similar form,

𝒮(l1,l2)=(S(l1,l2,l2)0000S(l1,l2,l21)0000S(l1,l2,1l2)0000S(l1,l2,l2)).
A simple analysis shows that there is a nonstrict selection rule of the principal quantum number nq, restricting the number of phonon modes that significantly contribute to the vibrational resonance (for given principal quantum numbers of electrons and holes, the nonstrictly allowed nqs are those for which the overlap integral in Eq. (4) is of the order of unity whereas the nonstrictly forbidden nqs are those for which the overlap integral is much less than unity). We further set nq = 1 and assume that all phonon modes of angular momenta lq = 0, 1, 2,... and momentum projections mq = −lq, −lq + 1,...,lq − 1, lq are degenerate in energy. Using the selection rules prescribed to the electron–phonon interaction by the Clebsch–Gordan coefficients in Eq. (3), it is easy to show that only phonon modes of 1+ min(l1, l2) angular momenta |l1l2|, |l1l2| + 2,..., l1 + l2 − 2, and l1 + l2 can resonantly interact with the electronic subsystem of the QD. It can be also shown that the number of phonon modes coupling the pair of hole states does not depend on the projection of l1 and is given by
k(l1,l2,m2)=ν=11+min(l1,l2)min(2l1+1,2lqν+1,2l1+2lqνl2|m2|+1).

Noticing that the interaction-induced energy shifts in the QD energy spectrum are usually relatively small, i.e. qh¯Ω|Φp;p(q)|2Ep in Eq. (11a), we consider energies p1(2) to be independent of the angular momentum projections m1 and m2. Since the diagonalization of the block Hamiltonian in Eq. (18) can be performed via diagonalizing each of its blocks separately, the problem of spectrum renormalization is reduced to 2l2 + 1 identical renormalization problems for the pair of nondegenerate states of the same energies coupled through k(l1, l2, m2) degenerate phonon modes. The block of the Hamiltonian (l1, l2) describing such coupling is

H(l1,l2,m2)H˜e,ph(k)=(E˜p2Vp2;p1(q1)Vp2;p1(q2)Vp2;p1(qk)Vp2;p1(q1)*E˜p1+h¯Ω00Vp2;p1(q2)*0E˜p1+h¯Ω0Vp2;p1(qk)*00E˜p1+h¯Ω).
The diagonal elements of this matrix are the energies of states |ψ̃1〉 = |p2〉 |0q〉, |ψ̃2〉 = |p1〉 |1q1〉, |ψ̃3〉 = |p1〉 |1q2〉,..., and |ψ̃k+1〉 = |p1〉 |1qk〉, where kets |0q〉 and |1qν〉 stand for the vacuum of phonon modes and one LO phonon in mode qν.

Some simple but tedious algebra yields the following unitary operator diagonalizing the Hamiltonian (k):

S(l1,l2,m2)S(k)=(S1;1(k)S1;2(k)0000S2;1(k)S2;2(k)S2;3(k)S2;4(k)S2;k(k)S2;k+1(k)S3;1(k)S3;2(k)S3;3(k)S3;4(k)S3;k(k)S3;k+1(k)S4;1(k)S4;2(k)0S4;4(k)S4;k(k)S4;k+1(k)Sk;1(k)Sk;2(k)00Sk;k(k)Sk;k+1(k)Sk+1;1(k)Sk+1;2(k)000Sk+1;k+1(k)),
where
S1;1(k)=(α+δ(k))χ(k)/β(k,+),S1;2(k)=(αδ(k))χ(k)/β(k,),
with α = p2p1Ω;
Sn;1(k)=2Vp2;p1(qn1)*χ(k)/β(k,+),Sn;2(k)=2Vp2;p1(qn1)*χ(k)/β(k,)
for 2 ≤ nk + 1;
Sn;m(k)=Vp2;p1(qm1)Vp2;p1(qn1)*/(V(m1)V(m2))
for 2 ≤ n < mk + 1;
Sn;n(k)=V(n2)/V(n1)
for 3 ≤ nk + 1;
β(k,±)=[(α±δ(k))2+4(V(k))2]1/2,χ(k)=|Vp2;p1(qk)|/Vp2;p1(qk)*,
δ(k)=[α2+4(V(k))2]1/2,V(k)=(ν=1k|Vp2;p1(qν)|2)1/2.
The energies and wave functions of the resulting polaron-like states are:
1,2(k)=12(E˜p1+E˜p2+h¯Ω±δ(k)),
3(k)=4(k)==k+1(k)=E˜p1+h¯Ω
and
|ψ^n=m=1k+1Sm;n(k)*|ψ˜m.

We illustrate the outlined renormalization procedure by explicitly writing out the matrices (l1, l2) and 𝒮 (l1, l2) for three simple cases. According to the selection rules and Eq. (20), the lower hole state with l1 = 0 is coupled to each of the three upper states with l2 = 1 and m2 = {−1, 0, 1} through one (k = 1) phonon mode with lq = 1 and mq = −m2. We thus find that

(0,1)=(E˜p2(+1)Vp2;p1(1,1,1)0000Vp2;p1(1,1,1)*E˜p1+h¯Ω000000E˜p2(0)Vp2;p1(1,1,0)0000Vp2;p1(1,1,0)*E˜p1+h¯Ω000000E˜p2(1)Vp2;p1(1,1,+1)0000Vp2;p1(1,1,+1)*E˜p1+h¯Ω)
and
𝒮(0,1)=(S(0,1,+1)000S(0,1,0)000S(0,1,1)),
where the superscript of E˜p2(m2) denotes (for the sake of clarity) the angular momentum projection of the upper state.

For l1 = 1 and l2 = 0 the selection rules give lq = 1 and mq = m2, so that the upper state is coupled to the lower states via three (k = 3) phonon modes, and we are led to the single-block matrices

(1,0)=(E˜p2Vp2;p1(1,1,+1)Vp2;p1(1,1,0)Vp2;p1(1,1,1)Vp2;p1(1,1,+1)*E˜p1(+1)+h¯Ω00Vp2;p1(1,1,0)*0E˜p1(0)+h¯Ω0Vp2;p1(1,1,1)*00E˜p1(1)+h¯Ω)
and 𝒮 (1, 0) = S(1, 0, 0).

The situation of states with l1 = 1 and l2 = 2 is a bit more intricate. These states are coupled through three phonon modes with lq1 = 1 and five modes with lq2 = 3. Both (1, 2) and 𝒮 (1, 2) have therefore five blocks each, with the blocks of the Hamiltonian matrix given by

H(1,2,±2)=(E˜p2(±2)Vp2;p1(1,3,3)Vp2;p1(1,3,2)Vp2;p1(1,3,1)Vp2;p1(1,1,1)Vp2;p1(1,3,3)*E˜p1(1)+h¯Ω000Vp2;p1(1,3,2)*0E˜p1(0)+h¯Ω00Vp2;p1(1,3,1)*00E˜p1(±1)+h¯Ω0Vp2;p1(1,1,1)*000E˜p1(±1)+h¯Ω),
H(1,2,±1)==(E˜p2(±1)Vp2;p1(1,3,2)Vp2;p1(1,3,1)Vp2;p1(1,3,0)Vp2;p1(1,1,1)Vp2;p1(1,1,0)Vp2;p1(1,3,2)*E˜p1(1)+h¯Ω0000Vp2;p1(1,3,1)*0E˜p1(0)+h¯Ω000Vp2;p1(1,3,0)*00E˜p1(±1)+h¯Ω00Vp2;p1(1,1,1)*000E˜p1(0)+h¯Ω0Vp2;p1(1,1,0)*0000E˜p1(±1)+h¯Ω)
and
H(1,2,0)==(E˜p2(0)Vp2;p1(1,3,1)Vp2;p1(1,3,0)Vp2;p1(1,3,1)Vp2;p1(1,1,1)Vp2;p1(1,1,0)Vp2;p1(1,1,1)Vp2;p1(1,3,1)*E˜p1(1)+h¯Ω00000Vp2;p1(1,3,0)*0E˜p1(0)+h¯Ω0000Vp2;p1(1,3,±1)*00E˜p1(±1)+h¯Ω000Vp2;p1(1,1,1)*000E˜p1(1)+h¯Ω00Vp2;p1(1,1,0)*0000E˜p1(0)+h¯Ω0Vp2;p1(1,1,±1)*00000E˜p1(±1)+h¯Ω).

Acknowledgments

The authors gratefully acknowledge the financial support from the Ministry of Education and Science of the Russian Federation (through its Grant No. 14.B25.31.0002) and the Russian Foundation for Basic Research (through its Grants Nos. 12-02-01263 and 12-02-00938). The Ministry of Education and Science of the Russian Federation also supports A.S.B. and M.Yu.L., through its scholarships of the President of the Russian Federation for young scientists and graduate students (2013–2015). A.S.B. gratefully acknowledges the Dynasty Foundation Support Program for Physicists. The work of I.D.R. is sponsored by the Australian Research Council, through its Discovery Early Career Researcher Award DE120100055.

References and links

1. G. Yu, Q. Liang, Y. Jia, and J. Dong, “Phonon sidebands of photoluminescence in single wall carbon nanotubes,” J. Appl. Phys. 107, 024314 (2010). [CrossRef]  

2. S. Sohal, Y. Alivov, Z. Fan, and M. Holtz, “Role of phonons in the optical properties of magnetron sputtered ZnO studied by resonance Raman and photoluminescence,” J. Appl. Phys. 108, 053507 (2010). [CrossRef]  

3. S. Y. Kruchinin, A. V. Fedorov, A. V. Baranov, T. S. Perova, and K. Berwick, “Double quantum dot photoluminescence mediated by incoherent reversible energy transport,” Phys. Rev. B 81, 245303 (2010). [CrossRef]  

4. C. H. Ahn, S. K. Mohanta, N. E. Lee, and H. K. Cho, “Enhanced exciton–phonon interactions in photoluminescence of ZnO nanopencils,” Appl. Phys. Lett. 94, 261904 (2009). [CrossRef]  

5. S. K. Tripathy, G. Xu, X. Mu, Y. J. Ding, M. Jamil, R. A. Arif, N. Tansu, and J. B. Khurgin, “Phonon–assisted ultraviolet anti-stokes photoluminescence from GaN film grown on Si (111) substrate,” Appl. Phys. Lett. 93, 201107 (2008). [CrossRef]  

6. T. Voss, C. Bekeny, L. Wischmeier, H. Gafsi, S. Borner, W. Schade, A. C. Mofor, A. Bakin, and A. Waag, “Influence of exciton–phonon coupling on the energy position of the near-band-edge photoluminescence of ZnO nanowires,” Appl. Phys. Lett. 89, 182107 (2006). [CrossRef]  

7. I. D. Rukhlenko and A. V. Fedorov, “Resonant photoluminescence of quantum dots: Dynamics of electronic subsystem,” Bull. Russ. Acad. Sci. 70, 126 (2006).

8. V. V. Ursaki, I. M. Tiginyanu, V. V. Zalamai, V. M. Masalov, E. N. Samarov, G. A. Emelchenko, and F. Briones, “Photoluminescence and resonant Raman scattering from ZnO-opal structures,” J. Appl. Phys. 96, 1001–1006 (2004). [CrossRef]  

9. T. S. Kim, Y.-H. Kil, H. D. Yang, J.-H. Yang, W.-K. Hong, S. Kang, T. S. Jeong, and K.-H. Shim, “Growth and characterization of Si1−xGex QDs on Si/Si0.8Ge0.2 layer,” Electron. Mater. Lett. 8, 559–563 (2012). [CrossRef]  

10. E. S. F. Neto, S. W. da Silva, P. C. Morais, M. I. Vasilevskiy, M. A. P. da Silva, and N. O. Dantas, “Resonant Raman scattering in CdSx Se1−x nanocrystals: Effects of phonon confinement, composition, and elastic strain,” J. Raman Spectrosc. 42, 1660–1669 (2011). [CrossRef]  

11. K. Raulin, S. Turrell, B. Capoen, C. Kinowski, V. T. T. Tran, M. Bouazaoui, and O. Cristini, “Raman characterization of localized CdS nanostructures synthesized by UV irradiation in sol-gel silica matrices,” J. Ram. Spectrosc. 42, 1366–1372 (2011). [CrossRef]  

12. S. Dhara, A. K. Arora, S. Bera, and J. Ghatak, “Multiphonon probe in ZnS quantum dots,” J. Raman Spectrosc. 41, 1102–1105 (2010). [CrossRef]  

13. S. K. Gupta, R. Desai, P. K. Jha, S. Sahoo, and D. Kirin, “Titanium dioxide synthesized using titanium chloride: Size effect study using Raman spectroscopy and photoluminescence,” J. Ram. Spectrosc. 41, 350–355 (2010).

14. A. V. Baranov, T. S. Perova, R. A. Moore, S. Solosin, A. V. Fedorov, V. Yam, V. Thanh, and D. Bouchier, “Polarized Raman spectroscopy of multilayer Ge/Si(001) quantum dot heterostructures,” J. Appl. Phys. 95, 2857–2863 (2004). [CrossRef]  

15. K. Rezgui, S. Aloulou, J. Rihani, and M. Oueslati, “Competition between strain and confinement effects on the crystalline quality of InAs/GaAs (001) quantum dots probed by Raman spectroscopy,” J. Raman Spectrosc. 43, 1964–1968 (2012). [CrossRef]  

16. A. Cros, N. Garro, A. Cantarero, J. Coraux, H. Renevier, and B. Daudin, “Raman scattering as a tool for the evaluation of strain in GaN/AlN quantum dots: The effect of capping,” Phys. Rev. B 76, 165403 (2007). [CrossRef]  

17. A. Cros, N. Garro, J. M. Llorens, A. García-Cristóbal, A. Cantarero, N. Gogneau, E. Monroy, and B. Daudin, “Raman study and theoretical calculations of strain in GaN quantum dot multilayers,” Phys. Rev. B 73, 115313 (2006). [CrossRef]  

18. A. V. Baranov, A. V. Fedorov, T. S. Perova, R. A. Moore, V. Yam, D. Bouchier, V. Thanh, and K. Berwick, “Analysis of strain and intermixing in single-layer Ge/Si quantum dots using polarized Raman spectroscopy,” Phys. Rev. B 73, 075322 (2006). [CrossRef]  

19. R. W. Meulenberg, T. Jennings, and G. F. Strouse, “Compressive and tensile stress in colloidal CdSe semiconductor quantum dots,” Phys. Rev. B 70, 235311 (2004). [CrossRef]  

20. S. Wageh, A. M. El-Nahas, A. A. Higazy, and M. A. M. Mahmoud, “Preparation and characterization of a novel system of CdS nanoparticles embedded in borophosphate glass matrix,” J. Alloys Comp. 555, 161–168 (2013). [CrossRef]  

21. G. H. Shih, C. G. Allen, and B. G. Potter Jr., “Interfacial effects on the optical behavior of Ge:ITO and Ge:NO nanocomposite films,” Nanotechnol. 23, 075203 (2012). [CrossRef]  

22. N. Tschirner, H. Lange, A. Schliwa, A. Biermann, C. Thomsen, K. Lambert, R. Gomes, and Z. Hens, “Interfacial alloying in CdSe/CdS heteronanocrystals: A Raman spectroscopy analysis,” Chem. Mater. 24, 311–318 (2012). [CrossRef]  

23. A. V. Baranov, Y. P. Rakovich, J. F. Donegan, T. S. Perova, R. A. Moore, D. V. Talapin, A. L. Rogach, Y. Masumoto, and I. Nabiev, “Effect of ZnS shell thickness on the phonon spectra in CdSe quantum dots,” Phys. Rev. B 68, 165306 (2003). [CrossRef]  

24. A. V. Fedorov, I. D. Rukhlenko, A. V. Baranov, and S. Y. Kruchinin, Optical Properties of Semiconductor Quantum Dots (Nauka, Saint Petersburg, 2011).

25. A. Rogach, ed., Semiconductor Nanocrystal Quantum Dots: Synthesis, Assembly, Spectroscopy and Applications (Springer, New York, 2008). [CrossRef]  

26. Y. Masumoto and T. Takagahara, eds., Semiconductor Quantum Dots: Physics, Spectroscopy and Applications (Springer, New York, 2002). [CrossRef]  

27. M. V. Mukhina, V. G. Maslov, A. V. Baranov, M. V. Artemyev, A. O. Orlova, and A. V. Fedorov, “Anisotropy of optical transitions in ordered ensemble of CdSe quantum rods,” Opt. Lett. 38, 3426–3428 (2013). [CrossRef]   [PubMed]  

28. A. S. Baimuratov, V. K. Turkov, I. D. Rukhlenko, and A. V. Fedorov, “Shape-induced anisotropy of intraband luminescence from a semiconductor nanocrystal,” Opt. Lett. 37, 4645–4647 (2012). [CrossRef]   [PubMed]  

29. A. S. Baimuratov, I. D. Rukhlenko, and A. V. Fedorov, “Engineering band structure in nanoscale quantum-dot supercrystals,” Opt. Lett. 38, 2259–2261 (2013). [CrossRef]   [PubMed]  

30. A. S. Baimuratov, I. D. Rukhlenko, V. K. Turkov, A. V. Baranov, and A. V. Fedorov, “Quantum-dot supercrystals for future nanophotonics,” Sci. Rep. 3, 1727 (2013). [CrossRef]  

31. V. Cesari, W. Langbein, and P. Borri, “Dephasing of excitons and multiexcitons in undoped and p-doped InAs/GaAs quantum dots-in-a-well,” Phys. Rev. B 82, 195314 (2010). [CrossRef]  

32. E. A. Muljarov and R. Zimmermann, “Exciton dephasing in quantum dots due to LO-phonon coupling: An exactly solvable model,” Phys. Rev. Lett. 98, 187401 (2007). [CrossRef]   [PubMed]  

33. K. Kojima and A. Tomita, “Influence of pure dephasing by phonons on exciton–photon interfaces: Quantum microscopic theory,” Phys. Rev. B 73, 195312 (2006). [CrossRef]  

34. A. Vagov, V. M. Axt, T. Kuhn, W. Langbein, P. Borri, and U. Woggon, “Nonmonotonous temperature dependence of the initial decoherence in quantum dots,” Phys. Rev. B 70, 201305 (2004). [CrossRef]  

35. R. R. Cooney, S. L. Sewall, E. A. Dias, D. M. Sagar, K. E. H. Anderson, and P. Kambhampati, “Unified picture of electron and hole relaxation pathways in semiconductor quantum dots,” Phys. Rev. B 75, 245311 (2007). [CrossRef]  

36. B. Patton, W. Langbein, U. Woggon, L. Maingault, and H. Mariette, “Time- and spectrally-resolved four-wave mixing in single CdTe/ZnTe quantum dots,” Phys. Rev. B 73, 235354 (2006). [CrossRef]  

37. M. R. Salvador, M. W. Graham, and G. D. Scholes, “Exciton–phonon coupling and disorder in the excited states of CdSe colloidal quantum dots,” J. Chem. Phys. 125, 184709 (2006). [CrossRef]  

38. S. Sanguinetti, E. Poliani, M. Bonfanti, M. Guzzi, E. Grilli, M. Gurioli, and N. Koguchi, “Electron–phonon interaction in individual strain-free GaAs/Al0.3Ga0.7 As quantum dots,” Phys. Rev. B 73, 125342 (2006). [CrossRef]  

39. D. Valerini, A. Cretí, M. Lomascolo, L. Manna, R. Cingolani, and M. Anni, “Temperature dependence of the photoluminescence properties of colloidal CdSe/ZnS core/shell quantum dots embedded in a polystyrene matrix,” Phys. Rev. B 71, 235409 (2005). [CrossRef]  

40. I. D. Rukhlenko, A. V. Fedorov, A. S. Baymuratov, and M. Premaratne, “Theory of quasi-elastic secondary emission from a quantum dot in the regime of vibrational resonance,” Opt. Express 19, 15461–15482 (2011). [CrossRef]  

41. A. V. Fedorov, A. V. Baranov, and K. Inoue, “Exciton–phonon coupling in semiconductor quantum dots: Resonant Raman scattering,” Phys. Rev. B 56, 7491 (1997). [CrossRef]  

42. A. V. Fedorov and A. V. Baranov, “Exciton–vibrational interaction of the frohlich type in quasi-zero-size systems,” J. Exp. Theor. Phys. 83, 610–618 (1996).

43. T. Itoh, M. Nishijima, A. I. Ekimov, C. Gourdon, A. L. Efros, and M. Rosen, “Polaron and exciton–phonon complexes in CuCl nanocrystals,” Phys. Rev. Lett. 74, 1645 (1995). [CrossRef]   [PubMed]  

44. I. D. Rukhlenko and A. V. Fedorov, “Penetration of electric fields induced by surface phonon modes into the layers of a semiconductor heterostructure,” Opt. Spectrosc. 101, 253–264 (2006). [CrossRef]  

45. I. D. Rukhlenko and A. V. Fedorov, “Propagation of electric fields induced by optical phonons in semiconductor heterostructures,” Opt. Spectrosc. 100, 238–244 (2006). [CrossRef]  

46. B. A. Carpenter, E. A. Zibik, M. L. Sadowski, L. R. Wilson, D. M. Whittaker, J. W. Cockburn, M. S. Skolnick, M. Potemski, M. J. Steer, and M. Hopkinson, “Intraband magnetospectroscopy of singly and doubly charged n-type self-assembled quantum dots,” Phys. Rev. B 74, 161302 (2006). [CrossRef]  

47. V. Preisler, R. Ferreira, S. Hameau, L. A. de Vaulchier, Y. Guldner, M. L. Sadowski, and A. Lemaitre, “Hole–LO-phonon interaction in InAs/GaAs quantum dots,” Phys. Rev. B 72, 115309 (2005). [CrossRef]  

48. J. Zhao, A. Kanno, M. Ikezawa, and Y. Masumoto, “Longitudinal optical phonons in the excited state of CuBr quantum dots,” Phys. Rev. B 68, 113305 (2003). [CrossRef]  

49. S. Hameau, J. N. Isaia, Y. Guldner, E. Deleporte, O. Verzelen, R. Ferreira, G. Bastard, J. Zeman, and J. M. Gerard, “Far-infrared magnetospectroscopy of polaron states in self-assembled InAs/GaAs quantum dots,” Phys. Rev. B 65, 085316 (2002). [CrossRef]  

50. A. V. Fedorov, A. V. Baranov, A. Itoh, and Y. Masumoto, “Renormalization of energy spectrum of quantum dots under vibrational resonance conditions,” Semiconductors 35, 1390–1397 (2001). [CrossRef]  

51. S. Hameau, Y. Guldner, O. Verzelen, R. Ferreira, G. Bastard, J. Zeman, A. Lemaitre, and J. M. Gerard, “Strong electron–phonon coupling regime in quantum dots: Evidence for everlasting resonant polarons,” Phys. Rev. Lett. 83, 4152 (1999). [CrossRef]  

52. S. Y. Kruchinin and A. V. Fedorov, “Renormalization of the energy spectrum of quantum dots under vibrational resonance conditions: Persistent hole burning spectroscopy,” Opt. Spectrosc. 100, 41–48 (2006). [CrossRef]  

53. R. P. Miranda, M. I. Vasilevskiy, and C. Trallero-Giner, “Nonperturbative approach to the calculation of multi-phonon Raman scattering in semiconductor quantum dots: Polaron effect,” Phys. Rev. B 74, 115317 (2006). [CrossRef]  

54. E. P. Pokatilov, S. N. Klimin, V. M. Fomin, J. T. Devreese, and F. W. Wise, “Multiphonon Raman scattering in semiconductor nanocrystals: Importance of nonadiabatic transitions,” Phys. Rev. B 65, 075316 (2002). [CrossRef]  

55. O. Verzelen, R. Ferreira, and G. Bastard, “Excitonic polarons in semiconductor quantum dots,” Phys. Rev. Lett. 88, 146803 (2002). [CrossRef]   [PubMed]  

56. T. Stauber, R. Zimmermann, and H. Castella, “Electron–phonon interaction in quantum dots: A solvable model,” Phys. Rev. B 62, 7336 (2000). [CrossRef]  

57. V. M. Fomin, V. N. Gladilin, J. T. Devreese, E. P. Pokatilov, S. N. Balaban, and S. N. Klimin, “Photoluminescence of spherical quantum dots,” Phys. Rev. B 57, 2415–2425 (1998). [CrossRef]  

58. T. Inoshita and H. Sakaki, “Density of states and phonon-induced relaxation of electrons in semiconductor quantum dots,” Phys. Rev. B 56, R4355 (1997). [CrossRef]  

59. I. B. Levinson and E. I. Rashba, “Threshold phenomena and bound states in the polaron problem,” Physics-Uspekhi 16, 892–912 (1974). [CrossRef]  

60. M. P. Chamberlain, C. Trallero-Giner, and M. Cardona, “Theory of one-phonon Raman scattering in semiconductor microcrystallites,” Phys. Rev. B 51, 1680–1693 (1995). [CrossRef]  

61. E. Evans and D. L. Mills, “Theory of inelastic scattering of slow electrons by long-wavelength surface optical phonons,” Phys. Rev. B 8, 4126–4139 (1973).

62. N. Mori and T. Ando, “Electron–optical-phonon interaction in single and double heterostructures,” Phys. Rev. B 40, 6175–6188 (1989). [CrossRef]  

63. A. V. Fedorov and A. V. Baranov, “Relaxation of charge carriers in quantum dots with the involvement of plasmon–phonon modes,” Semiconductors 38, 1101–1109 (2004). [CrossRef]  

64. O. Madelung, M. Schultz, and H. Weiss, eds., “Semiconductors. Physics of Group IV Elements and III–V Compounds,” Landolt-Börnstein, New Series, Group III, vol. 17, Pt. a (Springer-Verlag, Berlin, 1982).

65. K. Blum, Density Matrix Theory and Applications (Springer, Berlin, 2012). [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 (5)

Fig. 1
Fig. 1 Formation of polaron-like states from a pair of nondegenerate electron–hole states coupled via one, two, three or four degenerate LO phonon modes of energy Ω. The electron–phonon interaction transforms the degenerate state of the QD into two, three, four or five nondegenerate polaron-like states.
Fig. 2
Fig. 2 Feynman diagrams describing [(a) and (b)] creation and [(c) and (d)] annihilation of a polaron-like quasiparticle of energy p with the simultaneous (a) absorption of a photon, h̄ωL, and a phonon, Ω, (b) absorption of a photon and emission of a phonon, (c) emission of a photon, h̄ωλ, and absorption of a phonon, and (d) emission of a photon and a phonon.
Fig. 3
Fig. 3 Matrix elements and overlap integrals as functions of potential barrier heights for electrons and holes in a spherical InAs QD with R = 9.14 nm. The material parameters are: ε0 = 15.15, ε = 12.25, Eg = 418 meV, Ω = 29.5 meV, me = 0.0219m0, and mh = 0.43m0 [64] (m0 is the free-electron mass). For other parameter refer to the text.
Fig. 4
Fig. 4 Photoluminescence spectra of InAs QD with a pair of nondegenerate electronic states coupled through two phonon modes. The excitation frequencies are: h ¯ ω 2 ( 2 ) = 530.6 meV, h ¯ ω 3 ( 2 ) = 533.7 meV, and h ¯ ω 1 ( 2 ) = 536.8 meV; V(2) = 3.1 meV. For material parameters refer to the text.
Fig. 5
Fig. 5 Excitation photoluminescence spectra of InAs QD with a pair of nondegenerate electronic states coupled through two phonon modes. The detection frequencies are shown near the spectra. The parameter values and resonant frequencies ω j ( 2 ) (j = 1, 2, 3) are the same as in Fig. 4.

Equations (55)

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

H e , ph = p 1 , p 2 q V p 2 ; p 1 ( q ) a p 2 + a p 1 b q + H . c . ,
H e , L = p V p ; 0 ( L ) a p + + H . c .
H e , r = p , λ V p ; 0 ( λ ) a p + c λ + H . c . ,
F p = 2 δ l e , l h δ m e , m h 0 R n e , l e R n h , l h x 2 d x ,
φ p k 2 , p k 1 ( q ) = 4 ( 2 l q + 1 ) ( 2 l k 1 + 1 ) h ¯ Ω ( 2 l k 2 + 1 ) ε * R C l q 0 , l k 1 0 l k 2 0 C l q m q , l k 1 m k 1 l k 2 m k 2 n k 2 l k 2 , n k 1 l k 1 n q l q ,
n k 2 l k 2 , n k 1 l k 1 n q l q = 0 n q l q R n k 1 l k 1 R n k 2 l k 2 x 2 d x ,
R n k l k = 1 A n k l k × { κ l k ( α n k l k s ) j l k ( α n k l k d x ) , x 1 j l k ( α n k l k d ) κ l k ( α n k l k s x ) , x > 1 ,
A n k l k = κ l k + 1 ( α n k l k s ) κ l k 1 ( α n k l k s ) j l k 2 ( α n k l k d ) j l k + 1 ( α n k l k d ) j l k 1 ( α n k l k d ) κ l k 2 ( α n k l k s ) ,
α n k l k d = ( R / h ¯ ) 2 m k d E n k l k ,
α n k l k s = ( R / h ¯ ) 2 m k s ( V k E n k l k ) ;
m k s α n k l k d j l k ( α n k l k d ) j l k ( α n k l k d ) = m k d α n k l k s κ l k ( α n k l k s ) κ l k ( α n k l k s ) ,
n q l q = 1 ξ n q l q j l q + 1 ( ξ n q l q ) × { j l q ( ξ n q l q x ) , x 1 0 , x > 1 ,
F p F l e m e , l h m h n e n h = 2 δ l e , l h δ m e , m h A n e l e A n h l h ( κ l e ( α n e l e s ) κ l h ( α n h l h s ) 0 1 j l e ( α n e l e d x ) j l h ( α n h l h d x ) x 2 d x + j l e ( α n e l e d ) j l h ( α n h l h d ) 1 κ l e ( α n e l e s x ) κ l h ( α n h l h s x ) x 2 d x )
n k 2 l k 2 , n k 1 l k 1 n q l q = κ l k 1 ( α n k 1 l k 1 s ) κ l k 2 ( α n k 2 l k 2 s ) ξ n q l q j l q + 1 ( ξ n q l q ) A n k 1 l k 1 A n k 2 l k 2 × 0 1 j l q ( ξ n q l q x ) j l k 1 ( α n k 1 l k 1 d x ) j l k 2 ( α n k 2 l k 2 d x ) x 2 d x .
H ˜ = U + H U ,
U = exp ( p , q ( Φ p ; p ( q ) b q H . c . ) a p + a p )
E ˜ p = E p q h ¯ Ω | Φ p ; p ( q ) | 2
H ˜ e , ph = p 1 p 2 q V p 2 ; p 1 ( q ) a p 2 + a p 1 b q + H . c . ,
H ˜ e , L = p V p ; 0 ( L ) ( 1 + q ( Φ p ; p ( q ) b q H . c . ) ) a p + + H . c . ,
H ˜ e , r = p , λ V p ; 0 ( λ ) ( 1 + q ( Φ p ; p ( q ) b q H . c . ) ) a p + c λ + H . c .
H ^ e , L ( k ) = ( 0 0 0 H 1 ( L , k ) 0 0 0 H 2 ( L , k ) 0 0 0 H k + 1 ( L , k ) H 1 ( L , k ) * H 2 ( L , k ) * H k + 1 ( L , k ) * 0 ) ,
H 1 ( L , k ) = V p 2 ; 0 ( L ) S 1 ; 1 ( k ) + V p 1 ; 0 ( L ) ν = 1 k Φ p 1 ; p 1 ( q ν ) S 1 + ν ; 1 ( k ) ,
H 2 ( L , k ) = V p 2 ; 0 ( L ) S 1 ; 2 ( k ) + V p 1 ; 0 ( L ) ν = 1 k Φ p 1 ; p 1 ( q ν ) S 1 + ν ; 2 ( k ) ,
H n ( L , k ) = V p 1 ; 0 ( L ) ν = 1 k Φ p 1 ; p 1 ( q ν ) S 1 + ν ; n ( k )
H ^ e , r ( k ) = ( 0 0 0 H 1 , q 1 ( r , k ) H 1 , q 2 ( r , k ) H 1 , q k ( r , k ) 0 0 0 H 2 , q 1 ( r , k ) H 2 , q 2 ( r , k ) H 2 , q k ( r , k ) 0 0 0 H k + 1 , q 1 ( r , k ) H k + 1 , q 2 ( r , k ) H k + 1 , q k ( r , k ) H 1 , q 1 ( r , k ) * H 2 , q 1 ( r , k ) * H k + 1 , q 1 ( r , k ) * 0 0 0 H 1 , q 2 ( r , k ) * H 2 , q 1 ( r , k ) * H k + 1 , q 2 ( r , k ) * 0 0 0 H 1 , q k ( r , k ) * H 2 , q k ( r , k ) * H k + 1 , q k ( r , k ) * 0 0 0 ) ,
H 1 , q ( r , k ) = V p 1 ; 0 ( λ ) S 1 + q ; 1 ( k ) + V p 2 ; 0 ( λ ) Φ p 2 ; p 2 ( q ) S 1 ; 1 ( k ) ,
H 2 , q ( r , k ) = V p 1 ; 0 ( λ ) S 1 + q ; 2 ( k ) + V p 2 ; 0 ( λ ) Φ p 2 ; p 2 ( q ) S 1 ; 2 ( k ) ,
H n , q ( r , k ) = V p 1 ; 0 ( λ ) S 1 + q ; n ( k )
| i = | 0 | 0 q | 0 λ ,
| 1 = | 1 ( k ) | 0 q | 0 λ , , | k + 1 = | k + 1 ( k ) | 0 q | 0 λ ,
| f 1 = | 0 | 1 q 1 | 1 λ , , | f k = | 0 | 1 q k | 1 λ .
ρ μ ν ( t ) t = 1 i h ¯ [ H ^ ( t ) , ρ ( t ) ] μ ν γ μ ν ρ μ ν ( t ) + δ μ ν ν ν ζ ν ν ρ ν ν ( t ) ,
d 2 σ d Θ d ω λ = C ( ω λ ) η = 1 k ( μ = 1 k + 1 2 γ i μ γ μ μ 2 γ ^ i μ + γ ph Δ L μ 2 + γ i μ 2 | d μ ( L , k ) | 2 | d μ , q η ( λ , k ) | 2 Δ λ μ 2 + γ f μ 2 + ν = 2 k + 1 μ = 1 , ( μ 2 ) ν 1 2 ζ ν μ γ i ν γ μ μ γ ν ν 2 γ ν f Δ L μ 2 + γ i μ 2 | d μ ( L , k ) | 2 | d ν , q η ( λ , k ) | 2 Δ λ ν 2 + γ f ν 2 ) ,
d n ( L , k ) = F p 2 S 1 ; n ( k ) + F p 1 ν = 1 k Φ p 1 ; p 1 ( q ν ) S 1 + ν ; n ( k ) ,
d n , q ( λ , k ) = F p 1 S 1 + q ; n ( k ) + F p 2 Φ p 2 ; p 2 ( q ) S 1 ; n ( k ) .
( l 1 , l 2 ) = ( H ( l 1 , l 2 , l 2 ) 0 0 0 0 H ( l 1 , l 2 , l 2 1 ) 0 0 0 0 H ( l 1 , l 2 , 1 l 2 ) 0 0 0 0 H ( l 1 , l 2 , l 2 ) ) ,
𝒮 ( l 1 , l 2 ) = ( S ( l 1 , l 2 , l 2 ) 0 0 0 0 S ( l 1 , l 2 , l 2 1 ) 0 0 0 0 S ( l 1 , l 2 , 1 l 2 ) 0 0 0 0 S ( l 1 , l 2 , l 2 ) ) .
k ( l 1 , l 2 , m 2 ) = ν = 1 1 + min ( l 1 , l 2 ) min ( 2 l 1 + 1 , 2 l q ν + 1 , 2 l 1 + 2 l q ν l 2 | m 2 | + 1 ) .
H ( l 1 , l 2 , m 2 ) H ˜ e , ph ( k ) = ( E ˜ p 2 V p 2 ; p 1 ( q 1 ) V p 2 ; p 1 ( q 2 ) V p 2 ; p 1 ( q k ) V p 2 ; p 1 ( q 1 ) * E ˜ p 1 + h ¯ Ω 0 0 V p 2 ; p 1 ( q 2 ) * 0 E ˜ p 1 + h ¯ Ω 0 V p 2 ; p 1 ( q k ) * 0 0 E ˜ p 1 + h ¯ Ω ) .
S ( l 1 , l 2 , m 2 ) S ( k ) = ( S 1 ; 1 ( k ) S 1 ; 2 ( k ) 0 0 0 0 S 2 ; 1 ( k ) S 2 ; 2 ( k ) S 2 ; 3 ( k ) S 2 ; 4 ( k ) S 2 ; k ( k ) S 2 ; k + 1 ( k ) S 3 ; 1 ( k ) S 3 ; 2 ( k ) S 3 ; 3 ( k ) S 3 ; 4 ( k ) S 3 ; k ( k ) S 3 ; k + 1 ( k ) S 4 ; 1 ( k ) S 4 ; 2 ( k ) 0 S 4 ; 4 ( k ) S 4 ; k ( k ) S 4 ; k + 1 ( k ) S k ; 1 ( k ) S k ; 2 ( k ) 0 0 S k ; k ( k ) S k ; k + 1 ( k ) S k + 1 ; 1 ( k ) S k + 1 ; 2 ( k ) 0 0 0 S k + 1 ; k + 1 ( k ) ) ,
S 1 ; 1 ( k ) = ( α + δ ( k ) ) χ ( k ) / β ( k , + ) , S 1 ; 2 ( k ) = ( α δ ( k ) ) χ ( k ) / β ( k , ) ,
S n ; 1 ( k ) = 2 V p 2 ; p 1 ( q n 1 ) * χ ( k ) / β ( k , + ) , S n ; 2 ( k ) = 2 V p 2 ; p 1 ( q n 1 ) * χ ( k ) / β ( k , )
S n ; m ( k ) = V p 2 ; p 1 ( q m 1 ) V p 2 ; p 1 ( q n 1 ) * / ( V ( m 1 ) V ( m 2 ) )
S n ; n ( k ) = V ( n 2 ) / V ( n 1 )
β ( k , ± ) = [ ( α ± δ ( k ) ) 2 + 4 ( V ( k ) ) 2 ] 1 / 2 , χ ( k ) = | V p 2 ; p 1 ( q k ) | / V p 2 ; p 1 ( q k ) * ,
δ ( k ) = [ α 2 + 4 ( V ( k ) ) 2 ] 1 / 2 , V ( k ) = ( ν = 1 k | V p 2 ; p 1 ( q ν ) | 2 ) 1 / 2 .
1 , 2 ( k ) = 1 2 ( E ˜ p 1 + E ˜ p 2 + h ¯ Ω ± δ ( k ) ) ,
3 ( k ) = 4 ( k ) = = k + 1 ( k ) = E ˜ p 1 + h ¯ Ω
| ψ ^ n = m = 1 k + 1 S m ; n ( k ) * | ψ ˜ m .
( 0 , 1 ) = ( E ˜ p 2 ( + 1 ) V p 2 ; p 1 ( 1 , 1 , 1 ) 0 0 0 0 V p 2 ; p 1 ( 1 , 1 , 1 ) * E ˜ p 1 + h ¯ Ω 0 0 0 0 0 0 E ˜ p 2 ( 0 ) V p 2 ; p 1 ( 1 , 1 , 0 ) 0 0 0 0 V p 2 ; p 1 ( 1 , 1 , 0 ) * E ˜ p 1 + h ¯ Ω 0 0 0 0 0 0 E ˜ p 2 ( 1 ) V p 2 ; p 1 ( 1 , 1 , + 1 ) 0 0 0 0 V p 2 ; p 1 ( 1 , 1 , + 1 ) * E ˜ p 1 + h ¯ Ω )
𝒮 ( 0 , 1 ) = ( S ( 0 , 1 , + 1 ) 0 0 0 S ( 0 , 1 , 0 ) 0 0 0 S ( 0 , 1 , 1 ) ) ,
( 1 , 0 ) = ( E ˜ p 2 V p 2 ; p 1 ( 1 , 1 , + 1 ) V p 2 ; p 1 ( 1 , 1 , 0 ) V p 2 ; p 1 ( 1 , 1 , 1 ) V p 2 ; p 1 ( 1 , 1 , + 1 ) * E ˜ p 1 ( + 1 ) + h ¯ Ω 0 0 V p 2 ; p 1 ( 1 , 1 , 0 ) * 0 E ˜ p 1 ( 0 ) + h ¯ Ω 0 V p 2 ; p 1 ( 1 , 1 , 1 ) * 0 0 E ˜ p 1 ( 1 ) + h ¯ Ω )
H ( 1 , 2 , ± 2 ) = ( E ˜ p 2 ( ± 2 ) V p 2 ; p 1 ( 1 , 3 , 3 ) V p 2 ; p 1 ( 1 , 3 , 2 ) V p 2 ; p 1 ( 1 , 3 , 1 ) V p 2 ; p 1 ( 1 , 1 , 1 ) V p 2 ; p 1 ( 1 , 3 , 3 ) * E ˜ p 1 ( 1 ) + h ¯ Ω 0 0 0 V p 2 ; p 1 ( 1 , 3 , 2 ) * 0 E ˜ p 1 ( 0 ) + h ¯ Ω 0 0 V p 2 ; p 1 ( 1 , 3 , 1 ) * 0 0 E ˜ p 1 ( ± 1 ) + h ¯ Ω 0 V p 2 ; p 1 ( 1 , 1 , 1 ) * 0 0 0 E ˜ p 1 ( ± 1 ) + h ¯ Ω ) ,
H ( 1 , 2 , ± 1 ) = = ( E ˜ p 2 ( ± 1 ) V p 2 ; p 1 ( 1 , 3 , 2 ) V p 2 ; p 1 ( 1 , 3 , 1 ) V p 2 ; p 1 ( 1 , 3 , 0 ) V p 2 ; p 1 ( 1 , 1 , 1 ) V p 2 ; p 1 ( 1 , 1 , 0 ) V p 2 ; p 1 ( 1 , 3 , 2 ) * E ˜ p 1 ( 1 ) + h ¯ Ω 0 0 0 0 V p 2 ; p 1 ( 1 , 3 , 1 ) * 0 E ˜ p 1 ( 0 ) + h ¯ Ω 0 0 0 V p 2 ; p 1 ( 1 , 3 , 0 ) * 0 0 E ˜ p 1 ( ± 1 ) + h ¯ Ω 0 0 V p 2 ; p 1 ( 1 , 1 , 1 ) * 0 0 0 E ˜ p 1 ( 0 ) + h ¯ Ω 0 V p 2 ; p 1 ( 1 , 1 , 0 ) * 0 0 0 0 E ˜ p 1 ( ± 1 ) + h ¯ Ω )
H ( 1 , 2 , 0 ) = = ( E ˜ p 2 ( 0 ) V p 2 ; p 1 ( 1 , 3 , 1 ) V p 2 ; p 1 ( 1 , 3 , 0 ) V p 2 ; p 1 ( 1 , 3 , 1 ) V p 2 ; p 1 ( 1 , 1 , 1 ) V p 2 ; p 1 ( 1 , 1 , 0 ) V p 2 ; p 1 ( 1 , 1 , 1 ) V p 2 ; p 1 ( 1 , 3 , 1 ) * E ˜ p 1 ( 1 ) + h ¯ Ω 0 0 0 0 0 V p 2 ; p 1 ( 1 , 3 , 0 ) * 0 E ˜ p 1 ( 0 ) + h ¯ Ω 0 0 0 0 V p 2 ; p 1 ( 1 , 3 , ± 1 ) * 0 0 E ˜ p 1 ( ± 1 ) + h ¯ Ω 0 0 0 V p 2 ; p 1 ( 1 , 1 , 1 ) * 0 0 0 E ˜ p 1 ( 1 ) + h ¯ Ω 0 0 V p 2 ; p 1 ( 1 , 1 , 0 ) * 0 0 0 0 E ˜ p 1 ( 0 ) + h ¯ Ω 0 V p 2 ; p 1 ( 1 , 1 , ± 1 ) * 0 0 0 0 0 E ˜ p 1 ( ± 1 ) + h ¯ Ω ) .
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.