This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Nonlinear optical properties and Kerr nonlinearity of Rydberg excitons
in Cu2O quantum wells

David Ziemkiewicz [email protected]    Gerard Czajkowski    Karol Karpiński    Sylwia Zielińska-Raczyńska Institute of Mathematics and Physics, Technical University of Bydgoszcz,
Al. Prof. S. Kaliskiego 7, 85-789 Bydgoszcz, Poland
Abstract

The quantum confiment of Rydberg excitons (REs) in quantum structures opens the way towards considering nonlinear interactions in such systems. We present a theoretical calculation of optical functions in the case of a nonlinear coupling between REs in a quantum well with an electromagnetic wave. Using the Real Density Matrix Approach (RDMA), the analytical expressions for a linear and nonlinear absorption are derived and numerical calculations for Cu20 quantum wells are performed. The results indicate the conditions in which quantum well confinement states can be observed in linear and nonlinear optical spectra. The Kerr nonlinearity and self-phase modulation in such a system are studied. The effect of Rydberg blockade and the associated optical bleaching are also discussed and confronted with available experimental data.

pacs:
78.20.-e, 71.35.Cc, 71.36.+c

I Introduction

Rydberg physics in semiconductors has started in 2014 by an observation of highly excited excitonic states with principal quantum numbers as high as nn=25 in cuprous oxide, the material of a very large exciton binding energy [1]. This experiment revealed a plethora of Rydberg excitons’ unusual properties such as extraordinary large dimensions up to 1 μ\mum, long life-times of order of nanosecond, vulnerability to interactions with external fields and restrictions of their coupling arising from Rydberg blockade, which precludes a simultaneous excitation of two Rydberg excitons that are separated by less then a blockade radius rbr_{b}. A lot of papers have been devoted to studies of spectroscopic characteristic of REs in natural and synthetic bulk systems of Cu2O [2, 3, 4] (see more references therein). Simultaneously, the explorations of RE in the field of quantum optics have begun by demonstration of a generation and control of strong excitonic interactions with the help of two-color pump-probe technique [5], Rydberg exciton-assisted coupling between microwave and optical fields [6] and the experimental verification of the strong coupling of REs to cavity photons [7]. Moreover, some efforts have been made to investigate nonlinear interactions of REs with electromagnetic fields [8, 9]. The recent one-photon experiment has shown a giant nonlinear optical index in a bulk Cu2O crystal, caused by sharp Rydberg resonances and revealed a Kerr phase-shift much larger than in typical nonlinear crystals [10]. Interesting, giant microscopic dimensions of Rydberg excitons together with an intrinsic Rydberg blockade effect in cuprous oxide cause enhanced nonlinearities at much smaller densities compared with other semiconductors [1, 11].

Those results indicate that Rydberg excitons are a unique platform for obtaining strong interactions in solid systems and allow one to hope for a realization, in a close future, of solid state masers [12, 13] and few-photon devices. The first step to achieve a scalable solid-state platform characterized by controlled interactions between Rydberg excitons and photons to realize such technologically demanding miniaturized systems, consisted in an investigation of REs’ properties in strongly confined systems such as quantum dots, wires or wells [14, 15, 16, 17]. The experiment, which has verified a change of oscillator strength due to quantum confinement of REs in a nanoscale system [17], is an important step towards exploiting their large nonlinearities for quantum applications. The recent progress in fabricating synthetic cuprous oxide elements has shown an enormous progress of their quality manifested by observations of high excitonic states [4, 18, 19] and now the natural direction of subsequent explorations seems to be the study of a nonlinear interaction between confined REs and light. A great challenge in quantum optics is an accomplishment of the gigant Kerr nonlinearities in solid-state low-dimensional media. This phenomenon was realized in semiconductor quantum wells mostly under the conditions of the ellectromagnetically induced transparency [20, 21, 22] or in ultrathin gold films [23]. In our paper we propose a realization of the Kerr nonlinearity in the Cu2O quantum well with REs, taking advantage from the fact that confinement effects result in a significant optical Kerr susceptibility.

The theoretical tool which we use to calculate the optical functions for nonlinear interacion of electromagnetic radiation with Rydberg excitons in a quantum well is a mesoscopic method, called Real Density Matrix Approach [24, 25]. It allows for a calculation of transition probability amplitudes taking into account finite lifetimes, all mutual and external interactions and a system geometry. The detailed description of RDMA as well as a presentation of the iteration procedure, which allows one to obtain a nonlinear susceptibility for Rydberg excitons confined in a quantum well, is presented in Sections II and III. The effect of the Rydberg blockade is also included into our treatment and is considered in Sec. IV. Phase-sensitive Kerr nonlinearity appearing in discussed system is examined in Sec. V. Sec. VI contains the presentation of numerical results and their discussion, while the summary and conclusions of our paper are presented in the last Sec.VII.

II Real Density Matrix Approach

II.1 Basic equations

Our discussion follows the scheme of Refs.[8, 10] adapted for the case of a quantum well. In the RDMA approach that nonlinear response will be described by a set of three coupled constitutive equations: for the coherent amplitude Y(𝐫1,𝐫2)Y({\bf r}_{1},{\bf r}_{2}) representing the exciton density, for the density matrix C(𝐫1,𝐫2)C({\bf r}_{1},{\bf r}_{2}) for electrons (assuming a non-degenerate conduction band), and for the density matrix for the holes D(𝐫1,𝐫2)D({\bf r}_{1},{\bf r}_{2}) in the valence band. Denoting Y(𝐫1,𝐫2)=Y12Y({\bf r}_{1},{\bf r}_{2})=Y_{12}, the constitutive equations take the following form [16]
- for the coherent amplitude

itY12HQWY12=𝐌𝐄(𝐑12)\displaystyle{i}\hbar\partial_{t}Y_{12}-H_{\hbox{\tiny QW}}Y_{12}=-{\bf M}{\bf E}({\bf R}_{12})
+𝐄1𝐌0C12+𝐄2𝐌0D12+i(Y12t)irrev,\displaystyle+{\bf E}_{1}{\bf M}_{0}C_{12}+{\bf E}_{2}{\bf M}_{0}D_{12}+{i}\hbar\left(\frac{\partial Y_{12}}{\partial t}\right)_{{\rm irrev}}, (1)

- for the conduction band

itC12+HeeC12=𝐌0(𝐄1Y12𝐄2Y21)\displaystyle{i}\hbar\partial_{t}C_{12}+H_{ee}C_{12}={\bf M}_{0}({\bf E}_{1}Y_{12}-{\bf E}_{2}Y^{*}_{21})
+i(C12t)irrev,\displaystyle+{i}\hbar\left(\frac{\partial C_{12}}{\partial t}\right)_{{\rm irrev}}, (2)

- for the valence band

itD21HhhD21=𝐌0(𝐄2Y12𝐄1Y21)\displaystyle{i}\hbar\partial_{t}D_{21}-H_{hh}D_{21}={\bf M}_{0}({\bf E}_{2}Y_{12}-{\bf E}_{1}Y^{*}_{21})
+i(D21t)irrev,\displaystyle+{i}\hbar\left(\frac{\partial D_{21}}{\partial t}\right)_{{\rm irrev}}, (3)

where the operator HQWH_{\hbox{\tiny QW}} is the quantum well Hamiltonian, which includes the terms Ve,VhV_{e},V_{h} related to the electron and hole confinement and the mutual Coulomb interaction VehV_{eh}

HQW=Eg22meze222mhzh222MtotR2\displaystyle H_{\hbox{\tiny QW}}=E_{g}-\frac{\hbar^{2}}{2m_{e}}\partial_{z_{e}}^{2}-\frac{\hbar^{2}}{2m_{h}}\partial_{z_{h}}^{2}-\frac{\hbar^{2}}{2M_{\hbox{\tiny tot}}}\hbox{\boldmath$\nabla$}_{R_{\parallel}}^{2}
22μρ2+Ve(ze)+Vh(zh)+Veh,\displaystyle-\frac{\hbar^{2}}{2\mu}\hbox{\boldmath$\nabla$}^{2}_{\rho}+V_{e}(z_{e})+V_{h}(z_{h})+V_{eh}, (4)

with the separation of the center-of-mass coordinate 𝐑{\bf R}_{\parallel} from the relative coordinate 𝝆\rho on the plane (x,y)(x,y), e.g. ρ=(r1r2)\boldmath\rho=(r_{1}-r_{2})_{\parallel} and

Hee=22me(1222),\displaystyle H_{ee}=-\frac{\hbar^{2}}{2m_{e}}(\hbox{\boldmath$\nabla$}_{1}^{2}-\hbox{\boldmath$\nabla$}_{2}^{2}),
Hhh=22mh(1222),\displaystyle H_{hh}=-\frac{\hbar^{2}}{2m_{h}}(\hbox{\boldmath$\nabla$}_{1}^{2}-\hbox{\boldmath$\nabla$}_{2}^{2}), (5)

and 𝐄𝟏=𝐄(𝐫𝟏){\bf E_{1}}={\bf E(r_{1})}, 𝐄𝟐=𝐄(𝐫𝟐){\bf E_{2}}={\bf E(r_{2})}. In the case of a quantum well with thickness that is significantly smaller than the light wavelength, one can assume uniform field 𝐄𝟏=𝐄𝟐=𝐄{\bf E_{1}}={\bf E_{2}}={\bf E}. The center of mass coordinate is

𝐑=𝐑12=mh𝐫1+me𝐫2mh+me.{\bf R}={\bf R}_{12}=\frac{m_{h}{\bf r}_{1}+m_{e}{\bf r}_{2}}{m_{h}+m_{e}}. (6)

In the above formulas me,mhm_{e},m_{h} are the electron and the hole effective masses (the effective mass tensors in general), MtotM_{tot} is the total exciton mass and μ\mu the reduced mass of electron-hole pair. The smeared-out transition dipole density 𝐌(𝐫{\bf M}({\bf r}) is related to the bilocality of the amplitude Y12Y_{12} and describes the quantum coherence between the macroscopic electromagnetic field and the inter-band transitions (see, for example, Refs. [24, 25]); the detailed derivation of 𝐌(𝐫){\bf M}({\bf r}) is described in Ref. [26]. We assume that the carrier motion in the zz-direction is governed by the no-escape boundary conditions. With this assumptions, the QW Hamiltonian has the form

HQW=Eg+pze22me+pzh22mh+V(ze)+V(zh)\displaystyle H_{\hbox{\tiny QW}}=E_{g}+\frac{p_{z_{e}}^{2}}{2m_{e}}+\frac{p_{z_{h}}^{2}}{2m_{h}}+V(z_{e})+V(z_{h})
+HCoul(2D)(𝝆),\displaystyle+H_{\hbox{\tiny Coul}}^{(2D)}(\hbox{\boldmath$\rho$}), (7)

where

V(ze,h)=0for0zL,\displaystyle V(z_{e,h})=0\qquad\hbox{for}\quad 0\leq z\leq L,
V(ze,h)=forz<0,z>L,\displaystyle V(z_{e,h})=\infty\qquad\hbox{for}\quad z<0,\quad z>L, (8)

HCoul(2D)H_{\hbox{\tiny Coul}}^{(2D)} is the two-dimensional Coulomb Hamiltonian

HCoul(2D)(𝝆)=𝐩22μe24πϵ0ϵbρ.H_{\hbox{\tiny Coul}}^{(2D)}(\hbox{\boldmath$\rho$})=\frac{{\bf p}^{2}_{\parallel}}{2\mu_{\parallel}}-\frac{e^{2}}{4\pi\epsilon_{0}\epsilon_{b}\rho}. (9)

We consider here the strong confinement regime, where the confinement energy exceeds the Coulomb energy. The resulting coherent amplitude Y12Y_{12} determines the excitonic part of the polarization of the medium

𝐏(𝐑,t)=2d3rM(𝐫)ReY12(𝐑,𝐫,t)\displaystyle{\bf P}({\bf R},t)=2\int{\rm d}^{3}r\,\textbf{M}^{*}({\bf r})\hbox{Re}~{}Y_{12}({\bf R},{\bf r},t)
=d3rM(𝐫)[Y12(𝐑,𝐫,t)+c.c],\displaystyle=\int{\rm d}^{3}r\textbf{M}^{*}({\bf r})[Y_{12}({\bf R},{\bf r},t)+\hbox{c.c}], (10)

where 𝐫=𝐫1𝐫2{\bf r}={\bf r}_{1}-{\bf r}_{2} is the electron-hole relative coordinate. The linear optical functions are obtained by solving the interband equation (II.1) together with the corresponding Maxwell equation, where the polarization (II.1) acts as a source. Using the entire set of constitutive equations (II.1)-(II.1) one can the nonlinear optical functions. While a general solution of this problem seems to be inaccessible, but in specific situations such a solution can be found, i.e., if one assumes that the matrices Y,CY,C and DD can be expanded in powers of the electric field 𝐄{\bf E}, an iteration scheme can be used.

The relevant expansion of the polarization in powers of the field has the form

P(𝐤,ω)=ϵ0E(𝐤,ω)[χ(1)+χ(3)(ω,ω,ω)|E(𝐤,ω)|2+],P({\bf k},\omega)=\epsilon_{0}E({\bf k},\omega)\left[\chi^{(1)}+\chi^{(3)}(\omega,-\omega,\omega)|E({\bf k},\omega)|^{2}+\ldots\right], (11)

where χ(1)\chi^{(1)} and χ(3)\chi^{(3)} are the linear and the nonlinear parts of the susceptibility. Although the above equations apparently resemble those describing the nonlinear case of the bulk crystal with Rydberg excitons [8] we present full theoretical approach for the sake of completeness and it should be stressed that taking into account the confinement interaction significantly changes the results.

II.2 Iteration

We calculate the QW optical functions iteratively from the constitutive equations (II.1)-(II.1). The first step in the iteration consists of solving the equation (II.1)( skipping the second and the third terms in its r.h.s.) which we take in the form

itY12(1)HQWY12(1)=𝐌𝐄+i(Y12(1)t)irrev.{i}\hbar\partial_{t}Y^{(1)}_{12}-H_{\hbox{\tiny QW}}Y^{(1)}_{12}=-{\bf M}{\bf E}+{i}\hbar\left(\frac{\partial Y_{12}^{(1)}}{\partial t}\right)_{\hbox{\tiny irrev}}. (12)

It should be mentioned that we use the long-wave approximation, which allows to neglect the spatial distribution of the electromagnetic wave inside the quantum well.

For the irreversible part, assuming a relaxation time approximation one gets

(Y12(1)t)irrev=1T2Y12=ΓY12.\left(\frac{\partial Y_{12}^{(1)}}{\partial t}\right)_{\hbox{\tiny irrev}}=-\frac{1}{T_{2}}Y_{12}=\frac{-\Gamma}{\hbar}Y_{12}. (13)

with Γ=/T2{\mit\Gamma}=\hbar/T_{2} being a dissipation constant. Considering nonlinear effects the non-resonant parts of the coherent amplitude YY have to be taken into account; so for the electric field 𝐄{\bf E} in the medium of the form

𝐄=𝐄(𝐑,t)+𝐄(𝐑,t)=𝐄0ei(𝐤𝐑ωt)+𝐄0ei(𝐤𝐑ωt),{\bf E}={\bf E}({\bf R},t)+{\bf E}^{*}({\bf R},t)={\bf E}_{0}e^{{i}({\bf kR}-\omega t)}+{\bf E}_{0}e^{-{i}({\bf kR}-\omega t)}, (14)

equation (12) generates two equations: one for an amplitude Y(1)exp(iωt)Y_{-}^{(1)}\,\propto\exp(-{i}\omega t), and the second for the non-resonant part Y+(1)exp(iωt)Y^{(1)}_{+}\,\propto\exp({i}\omega t),

i(iω+1T2)Y12+(1)HehY12+(1)=𝐌𝐄(𝐑,t),\displaystyle{i}\hbar\left(i\omega+\frac{1}{T_{2}}\right)Y^{(1)}_{12+}-H_{eh}Y^{(1)}_{12+}=-{\bf M}{\bf E}^{*}({\bf R},t),
(15)
i(iω+1T2)Y12(1)HQWY12(1)=𝐌𝐄(𝐑,t).\displaystyle{i}\hbar\left(-{i}\omega+\frac{1}{T_{2}}\right)Y^{(1)}_{12-}-H_{\hbox{\tiny QW}}Y^{(1)}_{12-}=-{\bf M}{\bf E}({\bf R},t).

In what follows we consider only one component of both 𝐄{\bf E} and 𝐌{\bf M}. Similarly as in Ref. [8], we look for the solution in terms of eigenfunctions of the Hamiltonian HQWH_{\hbox{\tiny QW}}, which now contains the confinement terms, so these eigenfunctions have the following form

ΨjmNeNh(r,ϕ,ze,zh)=ψjm(r,ϕ)ψL,Ne(1D)(ze)ψL,Nh(1D)(zh)\Psi_{jmN_{e}N_{h}}(r,\phi,z_{e},z_{h})=\psi_{jm}(r,\phi)\psi^{(1D)}_{L,N_{e}}(z_{e})\psi^{(1D)}_{L,N_{h}}(z_{h}) (16)

where r=(r,ϕ)\textbf{r}=(r,\phi) is the two-dimensional space vector, ψjm\psi_{jm} are the normalized eigenfunctions of the 2-dimensional Coulomb Hamiltonian,

ψjm(r,ϕ)=Rjm(r)eimϕ2π,\displaystyle\psi_{jm}(r,\phi)=R_{jm}(r)\frac{e^{im\phi}}{\sqrt{2\pi}},
Rjm=Cjm(4κra)me2κjmr/aM(j,2|m|+1,4κjmra),\displaystyle R_{jm}=C_{jm}\left(4\kappa\frac{r}{a^{*}}\right)^{m}e^{-2\kappa_{jm}r/a^{*}}\,M\left(-j,2|m|+1,4\kappa_{jm}\frac{r}{a^{*}}\right),
κjm=11+2(j+|m|),\displaystyle\kappa_{jm}=\frac{1}{1+2(j+|m|)}, (17)
Cjm=1a4κjm3/21(2m)![(j+2m)!]1/2[j!]1/2,\displaystyle C_{jm}=\frac{1}{a^{*}}4\kappa_{jm}^{3/2}\frac{1}{(2m)!}\frac{[(j+2m)!]^{1/2}}{[j!]^{1/2}},

with the Kummer function [27] M(a,b,z)M(a,b,z) (the confluent hypergeometric function), and ψα,N(1D)(z)\psi^{(1D)}_{\alpha,N}(z) (N=0,1,…) are the quantum oscillator eigenfunctions of the Hamiltonian (II.1)

ψL,Ne(1D)(ze)=2Lcos[(2Ne1)πzeL],\displaystyle\psi_{L,N_{e}}^{(1D)}(z_{e})=\sqrt{\frac{2}{L}}\cos\left[(2N_{e}-1)\pi\frac{z_{e}}{L}\right],
ψL,Nh(1D)(zh)=2Lcos[(2Nh1)πzhL].\displaystyle\psi_{L,N_{h}}^{(1D)}(z_{h})=\sqrt{\frac{2}{L}}\cos\left[(2N_{h}-1)\pi\frac{z_{h}}{L}\right].

The role of the amplitude Y12(1)Y_{12}^{(1)} obtained in such a way is two-fold. First, substituted into Eq. (II.1) gives the linear excitonic polarization P(1)P^{(1)} and from the relation P(1)=ϵ0χ(1)EP^{(1)}=\epsilon_{0}\chi^{(1)}E we can calculate the mean effective linear susceptibility, which is given by the following expression

χ(1)(ω)=\displaystyle\chi^{(1)}\left(\omega\right)=
=χ0(1)(aL)0Nmaxj=0Jfj(2D)ETjNETjN2(ω+iΓj)2,\displaystyle=\chi^{(1)}_{0}\left(\frac{a^{*}}{L}\right)\,\sum\limits_{0}^{N_{\hbox{\tiny max}}}\sum\limits_{j=0}^{J}\frac{f_{j}^{(2D)}E_{TjN}}{E_{TjN}^{2}-(\hbar\omega+i{\mit\Gamma}_{j})^{2}}, (18)

where the summation is over the confinement state number NN and excitonic state number jj, where j=0j=0 is the lowest excitonic state. The oscillator strengths fj(2D)f_{j}^{(2D)} are given by

fj(2D)=48(j+1)(j+2)(j+3/2)51(1+2κjρ0)8\displaystyle f_{j}^{(2D)}=48\frac{(j+1)(j+2)}{\left(j+3/2\right)^{5}}\frac{1}{(1+2\kappa_{j}\rho_{0})^{8}}
×[F(j,4;3;4κjρ01+2κjρ0)]2,\displaystyle\times\left[F\left(-j,4;3;\frac{4\kappa_{j}\rho_{0}}{1+2\kappa_{j}\rho_{0}}\right)\right]^{2}, (19)
κj=12j+3,\displaystyle\kappa_{j}=\frac{1}{2j+3},

and the energy terms, including exciton binding energy EjE_{j} and quantum well contribution WNW_{N} are as follows

ETjN=Eg+WN+Ej,\displaystyle E_{TjN}=E_{g}+W_{N}+E_{j},
WN=(NπaL)2R,N=1,2,,\displaystyle W_{N}=\left(\frac{N\pi a^{*}}{L}\right)^{2}R^{*},\qquad N=1,2,\ldots, (20)
Ej=1(j+1δ)2R,j=0,1,2,\displaystyle E_{j}=-\frac{1}{(j+1-\delta)^{2}}R^{*},\quad j=0,1,2,\ldots

aa^{*} is the effective exciton Bohr radius, RR^{*} the exciton Rydberg energy, ρ0=r0/a\rho_{0}=r_{0}/a^{*} defines the coherence radius, and F(a,b;c,z)F(a,b;c,z) is the hypergeometric series [27]. The δ=0.2\delta=0.2 is the so-called quantum defect [1]; it should be mentioned that while the most common value of δ\delta is used here, smaller ones provide a better fit to many experimental results, especially at elevated temperatures [28]. The constant factor χ0(1)\chi^{(1)}_{0} has the form

χ0(1)=ϵbe4ρ0ΔLT,\chi_{0}^{(1)}=\epsilon_{b}e^{-4\rho_{0}}\Delta_{LT}, (21)

For simplicity, we can use only one confinement state number by considering only the largest contribution from Ne=Nh=NN_{e}=N_{h}=N. Due to the long wave approximation the validity of our considerations is limited regarding the quantum well width LL, which in turn entails the restriction of the highest observable confinement states NmaxN_{max}. Specifically, in the case of a thin quantum well, the considerable confinement energy WNW_{N} means that for higher NN and jj, the total energy ETjNE_{TjN} approaches the band gap, where higher absorption precludes the observation of confinement states.

III Iteration procedure: second step

Again, in order to present the detailed derivation of nonlinear susceptibility for a quantum well with Rydberg excitons we recall the procedure in general similar to that presented in [8], but considering here the low dimensional systems significantly changes the final results. Let us first consider a wave linearly polarized in the zz direction. Then Y±(1)Y_{\pm}^{(1)} (II.2) are inserted into the source terms of the conduction-band and valence band equations (II.1 - II.1). Solving for stationary solutions and making the long wave approximation, we obtain for both source terms

JC=M0ρ0(E1Y12(1)E2Y21(1))\displaystyle J_{C}=M_{0}\rho_{0}\left(E_{1}Y_{12}^{(1)}-E_{2}Y_{21}^{(1)*}\right)
=2iM0ρ0E02[Img(ω,𝐫)+Img(ω,𝐫)]=JV,\displaystyle=\frac{2{i}M_{0}\rho_{0}{E}_{0}^{2}}{\hbar}\left[\,\hbox{Im}\,g(-\omega,{\bf r})+\,\hbox{Im}\,g(\omega,{\bf r})\right]=J_{V}, (22)

where

g(±ω,𝐫)=jcjmNeNhΦjmNeNh(𝐫)ΩjmNeNhωi/T2jm.g(\pm\omega,{\bf r})=\sum_{j}\frac{c_{jmN_{e}N_{h}}\Phi_{jmN_{e}N_{h}}({\bf r})}{\Omega_{jmN_{e}N_{h}}\mp\omega-{i}/T_{2jm}}. (23)

If irreversible terms are well defined, the equations (II.1) can be solved and their solutions are then used in the saturating terms on the r.h.s. of the equations (II.1). Again as in the previous section, we will use a relaxation time approximation and the equations for the matrices CC and DD are as follows

(Ct)irrev\displaystyle\left(\frac{\partial C}{\partial t}\right)_{{\rm irrev}}
=1τ[C(𝐗,𝐫,t)f0e(𝐫)C(𝐗,𝐫=r0,t)]C(r0)T1,\displaystyle=-\frac{1}{\tau}\left[C({\bf X},{\bf r},t)-f_{0e}({\bf r})C({\bf X},{\bf r}=\textbf{r}_{0},t)\right]-\frac{C(r_{0})}{T_{1}},
(24)
(Dt)irrev\displaystyle\left(\frac{\partial D}{\partial t}\right)_{{\rm irrev}}
=1τ[D(𝐗,𝐫,t)f0h(𝐫)D(𝐗,𝐫=r0,t)]D(r0)T1,\displaystyle=-\frac{1}{\tau}\left[D({\bf X},{\bf r},t)-f_{0h}({\bf r})D({\bf X},{\bf r}=\textbf{r}_{0},t)\right]-\frac{D(r_{0})}{T_{1}},

where

𝐗=12(𝐫e+𝐫h),{\bf X}=\frac{1}{2}\left({\bf r}_{e}+{\bf r}_{h}\right), (25)

and f0e,f0hf_{0e},f_{0h} are normalized Boltzmann distributions for electrons and holes, respectively. The relaxation parameter T1T_{1} is due to interband recombination [29] and τ=1/Γj<<T1\tau=1/\Gamma_{j}<<T_{1} is the lifetime corresponding to radiative recombination. The functions C,DC,D must have the same pp-symmetry as the amplitudes YY. Thus we use the transport current density

jn(r)=i2me(12)|r1=r2=r,\textbf{j}_{n}(\textbf{r})=\frac{i\hbar}{2m_{e}}\left(\hbox{\boldmath$\nabla$}_{1}-\hbox{\boldmath$\nabla$}_{2}\right)|_{\hbox{\tiny{r}}_{1}=\hbox{\tiny{r}}_{2}=\textbf{r}}, (26)

and take the xx-component, which leads to the following expression for the modified distribution for electrons

f~0e(𝐫)=d3qqxf0e(q)eiqr\displaystyle\tilde{f}_{0e}({\bf r})=\int d^{3}q\,q_{x}\,f_{0e}(\textbf{q})\,e^{-i\textbf{qr}} (27)

with

f0e(𝐪)=2π(22πkB𝒯)2exp(2q22mekB𝒯),f_{0e}({\bf q})=\sqrt{2\pi}\left(\frac{\hbar^{2}}{2\pi k_{B}{\mathcal{T}}}\right)^{2}\exp\left(-\frac{\hbar^{2}q^{2}}{2m_{e}{k_{B}{\mathcal{T}}}}\right), (28)

where 𝒯\mathcal{T} is the temperature and kBk_{B} is the Boltzmann constant. The integral (27) can be evaluated analytically yielding

f~0e(r)=f0e~(ρ,ze,zh,ϕ)=π2rλth e\displaystyle\tilde{f}_{0e}(\textbf{r})=\tilde{f_{0e}}(\rho,z_{e},z_{h},\phi)=\sqrt{\frac{\pi}{2}}\frac{r}{\lambda_{\hbox{\tiny th e}}}
×[Φ1(ϕ)+Φ1(ϕ)]exp(r2+(zezh)22mekB𝒯2)\displaystyle\times[\Phi_{1}(\phi)+\Phi_{-1}(\phi)]\exp\left(-\frac{r^{2}+(z_{e}-z_{h})^{2}}{2}\frac{m_{e}k_{B}{\mathcal{T}}}{\hbar^{2}}\right)
=f~0e(ze,zh)f~0e(r,ϕ),\displaystyle=\tilde{f}_{0e}^{\perp}(z_{e},z_{h})\tilde{f}^{\parallel}_{0e}(r,\phi),
f~0e(ze,zh)=exp((zezh)22mekB𝒯2),\displaystyle\tilde{f}_{0e}^{\perp}(z_{e},z_{h})=\exp\left(-\frac{(z_{e}-z_{h})^{2}}{2}\frac{m_{e}k_{B}{\mathcal{T}}}{\hbar^{2}}\right), (29)
f~0e(r,ϕ)=π2rλth eexp(r22mekB𝒯2)[Φ1(ϕ)+Φ1(ϕ)],\displaystyle\tilde{f}^{\parallel}_{0e}(r,\phi)=\sqrt{\frac{\pi}{2}}\frac{r}{\lambda_{\hbox{\tiny th e}}}\exp\left(-\frac{r^{2}}{2}\frac{m_{e}k_{B}{\mathcal{T}}}{\hbar^{2}}\right)[\Phi_{1}(\phi)+\Phi_{-1}(\phi)],
r=x2+y2,\displaystyle r=\sqrt{x^{2}+y^{2}},

where

Φm(ϕ)=eimϕ2π,\displaystyle\Phi_{m}(\phi)=\frac{e^{im\phi}}{\sqrt{2\pi}}, (30)

and

λth e=(2mekB𝒯)1/2=2μmeRkB𝒯a,\displaystyle\lambda_{\hbox{\tiny th e}}=\left(\frac{\hbar^{2}}{m_{e}k_{B}{\mathcal{T}}}\right)^{1/2}=\sqrt{\frac{2\mu}{m_{e}}}\sqrt{\frac{R^{*}}{k_{B}{\mathcal{T}}}}a^{*}, (31)

is the so-called thermal length (here for electrons). Similarly, for the hole equilibrium distribution we have

f~0h(r)=f0h~(ρ,ze,zh,ϕ)=π2rλth h\displaystyle\tilde{f}_{0h}(\textbf{r})=\tilde{f_{0h}}(\rho,z_{e},z_{h},\phi)=\sqrt{\frac{\pi}{2}}\frac{r}{\lambda_{\hbox{\tiny th h}}}
×[Φ1(ϕ)+Φ1(ϕ)]exp(r2+(zezh)22mhkB𝒯2)=\displaystyle\times[\Phi_{1}(\phi)+\Phi_{-1}(\phi)]\exp\left(-\frac{r^{2}+(z_{e}-z_{h})^{2}}{2}\frac{m_{h}k_{B}{\mathcal{T}}}{\hbar^{2}}\right)=
=f~0h(ze,zh)f~0h(r,ϕ),\displaystyle=\tilde{f}_{0h}^{\perp}(z_{e},z_{h})\tilde{f}^{\parallel}_{0h}(r,\phi), (32)
f~0h(ze,zh)=exp((zezh)22mhkB𝒯2),\displaystyle\tilde{f}_{0h}^{\perp}(z_{e},z_{h})=\exp\left(-\frac{(z_{e}-z_{h})^{2}}{2}\frac{m_{h}k_{B}{\mathcal{T}}}{\hbar^{2}}\right),
f~0h(r,ϕ)=π2rλth hexp(r22mhkB𝒯2)[Φ1(ϕ)+Φ1(ϕ)],\displaystyle\tilde{f}^{\parallel}_{0h}(r,\phi)=\sqrt{\frac{\pi}{2}}\frac{r}{\lambda_{\hbox{\tiny th h}}}\exp\left(-\frac{r^{2}}{2}\frac{m_{h}k_{B}{\mathcal{T}}}{\hbar^{2}}\right)[\Phi_{1}(\phi)+\Phi_{-1}(\phi)],

with the hole thermal length

λth h=(2mhkB𝒯)1/2=2μmhRkB𝒯a.\displaystyle\lambda_{\hbox{\tiny th h}}=\left(\frac{\hbar^{2}}{m_{h}k_{B}{\mathcal{T}}}\right)^{1/2}=\sqrt{\frac{2\mu}{m_{h}}}\sqrt{\frac{R^{*}}{k_{B}{\mathcal{T}}}}a^{*}.

The matrices CC and DD are temperature-dependent, so they also can be used as an additional contribution for interpretation of temperature variations of excitonic optical spectra. However, the temperature dependence of relaxation constants Γn\Gamma_{n} remains a dominant mechanism influencing the spectra. Further, we will assume that our medium is excited homogeneously in 𝐗{\bf X} space. For pp excitons the matrices CC and DD relax to their values at r=r0r=r_{0}. In Cu2O, the dipole density can be approximated by M(r)rδ(rr0)\textbf{M}(\textbf{r})\propto\textbf{r}\delta(r-r_{0})[24], which leads to the following expressions for the matrices C,DC,~{}D

C(𝐫)=i[τJC(𝐫)τJC(r0)+T1f0e(𝐫)JC(r0)],\displaystyle C({\bf r})=-\frac{i}{\hbar}\left[\tau J_{C}({\bf r})-\tau J_{C}(r_{0})+T_{1}f_{0e}({\bf r})J_{C}(r_{0})\right],
(33)
D(𝐫)=i[τJV(𝐫)τJV(r0)+T1f0hH(𝐫)JV(r0)].\displaystyle D({\bf r})=-\frac{i}{\hbar}\left[\tau J_{V}({\bf r})-\tau J_{V}(r_{0})+T_{1}f_{0hH}({\bf r})J_{V}(r_{0})\right].

With the above expression the equation for the third order coherent amplitude Y12(3)Y^{(3)}_{12} takes the form

(ω+iT2)Y12(3)HQWY12(3)\displaystyle\hbar\left(\omega+\frac{\rm i}{T_{2}}\right)Y^{(3)}_{12-}-H_{QW}Y^{(3)}_{12-}
=M0ρ0(E1C12+E2D21)=E(𝐑,t)J~,\displaystyle=M_{0}\rho_{0}({E}_{1}C_{12}+{E}_{2}D_{21})={E}({\bf R},t)\tilde{J}_{-},
(ω+iT2)Y12+(3)HQWY12+(3)\displaystyle\hbar\left(-\omega+\frac{\rm i}{T_{2}}\right)Y^{(3)}_{12+}-H_{QW}Y^{(3)}_{12+}
=M0ρ0(E1C12+E2D21)=E(𝐑,t)J~+.\displaystyle=M_{0}\rho_{0}({E}^{*}_{1}C_{12}+{E}^{*}_{2}D_{21})={E}^{*}({\bf R},t)\tilde{J}_{+}. (34)

To define the source terms J~±\tilde{J}_{\pm} we use the fact that for most semiconductors T1τT_{1}\gg\tau. Therefore we retain only the terms proportional to T1T_{1}, obtaining

J~=iT1M0ρ0{JC(r0)f~0e(𝐫)\displaystyle\tilde{J}_{-}=-\frac{i}{\hbar}T_{1}M_{0}\rho_{0}\biggl{\{}J_{C}(r_{0})\tilde{f}_{0e}({\bf r})
+JV(r0)f~0h(𝐫)},\displaystyle+J_{V}(r_{0})\tilde{f}_{0h}({\bf r})\biggr{\}},
(35)
J~+=iM0ρ0T1{JC(r0)f~0e(𝐫+JV(r0)f~0h(𝐫)}.\displaystyle\tilde{J}_{+}=-\frac{i}{\hbar}M_{0}\rho_{0}T_{1}\biggl{\{}J_{C}(r_{0})\tilde{f}_{0e}({\bf r}+J_{V}(r_{0})\tilde{f}_{0h}({\bf r})\biggr{\}}.

From Y(3)Y^{(3)} one finds the third order polarization according to

P(3)(𝐑)=2d3rReM(𝐫)Y(3)(𝐑,𝐫)\displaystyle P^{(3)}({\bf R})=2\int{\rm d}^{3}{r}\hbox{Re}\,M({\bf r})Y^{(3)}({\bf R},{\bf r})
=d3rM(𝐫)(Y12(3)+Y12+(3)).\displaystyle=\int{\rm d}^{3}{r}\,M({\bf r})\left(Y^{(3)}_{12-}+Y^{(3*)}_{12+}\right). (36)

As in the case of linear amplitudes Y(1)Y^{(1)}, we expand the nonlinear amplitudes in terms of the eigenfunctions ΨmNeNh(𝐫)\Psi_{\ell mN_{e}N_{h}}({\bf r}).

The next application of the amplitude Y12(1)Y_{12}^{(1)} is related to the iteration process. Inserting Y12(1)Y_{12}^{(1)} in the source terms on the r.h.s. of Eqs. (II.1,II.1) and using appropriate expressions for the irreversible terms, one obtains the matrices C(2),D(2)C^{(2)},D^{(2)}, where the superscript indicates the order with respect to the electric field strength EE. Substituting the matrices into the saturating terms on the r.h.s. of Eq. (II.1) one obtains the equation for the nonlinear amplitude Y(3)Y^{(3)} which, with respect to Eq. (II.1), defines the nonlinear susceptibility χ(3)\chi^{(3)}. We obtain the following expression

χ(3)(ω)=(aL)χ0(3)\displaystyle\chi^{(3)}(\omega)=-\left(\frac{a^{*}}{L}\right)\chi_{0}^{(3)} (37)
×j,,NΓjjNETN[(ETjNω)2+Γj2][ETN2(E+iΓ)2],\displaystyle\times\sum\limits_{j,\ell,N}\frac{{\mit\Gamma}_{j}{\mathcal{F}}_{j\ell N}E_{T\ell N}}{[(E_{TjN}-\hbar\omega)^{2}+{\mit\Gamma_{j}}^{2}][E_{T\ell N}^{2}-(E+i{\mit\Gamma}_{\ell})^{2}]},

where =0,1,2\ell=0,1,2... The nonlinear oscillator strengths can be written as

jN=M(j,3,4κjρ0)M(,3,4κρ0)(1+2κρ0)4(1+2κjρ0)4\displaystyle{\mathcal{F}}_{j\ell N}=\frac{M(-j,3,4\kappa_{j}\rho_{0})\,M\left(-\ell,3,4\kappa_{\ell}\rho_{0}\right)}{(1+2\kappa_{\ell}\rho_{0})^{4}(1+2\kappa_{j}\rho_{0})^{4}}
×F(j,4;3;4κjρ01+2κjρ0)F(,4;3;4κ1ρ01+2κρ0)\displaystyle\times\,F\left(-j,4;3;\frac{4\kappa_{j}\rho_{0}}{1+2\kappa_{j}\rho_{0}}\right)\,F\left(-\ell,4;3;\frac{4\kappa_{\ell 1}\rho_{0}}{1+2\kappa_{\ell}\rho_{0}}\right)
×(j+1)(j+2)(2j+3)5(+1)(+2)(2+3)2[𝒜(22+3)βVNN(e)\displaystyle\times\frac{(j+1)(j+2)}{(2j+3)^{5}}\,\frac{(\ell+1)(\ell+2)}{(2\ell+3)^{2}}\left[{\mathcal{A}}\,\left(\frac{2}{2\ell+3}\right)^{\beta}\,V_{NN}^{(e)}\right.
+(22+3)γVNN(h)],\displaystyle\left.+{\mathcal{B}}\,\left(\frac{2}{2\ell+3}\right)^{\gamma}\,V_{NN}^{(h)}\right],
κj=12j+3,κ=12+3,\displaystyle\kappa_{j}=\frac{1}{2j+3},\quad\kappa_{\ell}=\frac{1}{2\ell+3}, (38)

where

χ0(3)=ϵ0(ϵbΔLT)2a3e4ρ0(1Γ1)\chi_{0}^{(3)}=\epsilon_{0}(\epsilon_{b}\Delta_{LT})^{2}a^{*3}e^{-4\rho_{0}}\left(\frac{1}{{\mit\Gamma}_{1}}\right) (39)

and the derivation of constants 𝒜\mathcal{A}, \mathcal{B} is presented in Appendix A. The potentials VNeNhV_{N_{e}N_{h}} are given by

VNeNh(e)=11𝑑x11𝑑ycos[(2Ne1)π2x]cos[(2Nh1)π2y]\displaystyle V_{N_{e}N_{h}}^{(e)}=\int\limits_{-1}^{1}dx\,\int\limits_{-1}^{1}\,dy\cos\left[\frac{(2N_{e}-1)\pi}{2}x\right]\cos\left[\frac{(2N_{h}-1)\pi}{2}y\right]
×exp[(L28a2λ~th e2)(xy)2]\displaystyle\times\exp\left[-\left(\frac{L^{2}}{8a^{*2}{\tilde{\lambda}_{\hbox{\tiny th e}}^{2}}}\right)(x-y)^{2}\right]
VNeNh(h)=11𝑑x11𝑑ycos[(2Ne1)π2x]cos[(2Nh1)π2y]\displaystyle V_{N_{e}N_{h}}^{(h)}=\int\limits_{-1}^{1}dx\,\int\limits_{-1}^{1}\,dy\cos\left[\frac{(2N_{e}-1)\pi}{2}x\right]\cos\left[\frac{(2N_{h}-1)\pi}{2}y\right]
×exp[(L28a2λ~th h2)(xy)2],\displaystyle\times\exp\left[-\left(\frac{L^{2}}{8a^{*2}{\tilde{\lambda}_{\hbox{\tiny th h}}^{2}}}\right)(x-y)^{2}\right], (40)

with the thermal lengths λ~th e,h\tilde{\lambda}_{\hbox{\tiny th e,h}} defined above in Eq. (31).

More detailed calculations of χ(3)\chi^{(3)} are presented in Appendix B and the table of material parameters is included in Appendix C. The long-wave approximation we have used in above calculations limits our results, which are appropriate for quantum wells thicknesses LL significantly bellow 1μm1\mu m.

IV Rydberg blockade

One of the important characteristics of the theoretical approach described above is the fact that it is derived under the assumption of a relatively low power level, when the medium is not saturated with excitons. Thus, the so-called Rydberg blockade [1] is not inherently present in the calculations and its effects have to be taken into account in a separate step. This has been done in [13, 10] and the description outlined below is an extension of the approaches presented in the cited works.

For an exciton with principal quantum number (j+1)(j+1), the blockade volume is given by [1]

VB=3107(j+1)7μm3.V_{B}=3\cdot 10^{-7}(j+1)^{7}~{}\mu m^{3}. (41)

Similarly to the recent experiments [10], we assume that the laser beam illuminating the sample has a circular beam spot of area SS of 0.1 mm2 and the sample length is LL; the volume , where the light can be absorbed and an exciton created is V=LSV=LS. Within this volume, a new exciton can be formed only when its location is outside of the blockade volume of existing excitons. Thus, assuming that the blockade volume is spherical, the upper limit of exciton density is the perfect sphere packing, where approximately 74%74\% of the volume is occupied, e. g. for the number of excitons NeN_{e}, NeVBV0.74N_{e}\frac{V_{B}}{V}\approx 0.74. However, the positions of the excitons formed within the laser beam are random and thus highly unlikely to form a perfect sphere packing. To estimate the practical upper limit of exciton density imposed by Rydberg blockade, a Monte Carlo simulation has been performed; within given volume VV, excitons with their associated blockade volumes are added at random positions and the number of attempts to place an exciton in a free space (not occupied by blockade volume) is counted. Then, the probability of excitation (inverse of the number of attempts) is calculated. The results are shown on the Fig. 1.

Refer to caption
Figure 1: The probability of excitation as a function of the volume occupied by Rydberg blockade.

One can see that the system is effectively saturated when the fraction of occupied volume approaches 0.2. An exponential function can be fitted to the data (dashed line), providing a simple model of saturation; the probability of excitation is

peexp(23.8[NeVBV]1.29).p_{e}\approx\exp\left(-23.8\left[\frac{N_{e}V_{B}}{V}\right]^{1.29}\right). (42)

When calculating the susceptibility from Eq. (II.2) and Eq. (37), one has to multiply the oscillator strengths FF by the above probability. This is a similar approach to that one used in [10] and [13], where also an exponential function exp(ANeVB/V)\exp(-AN_{e}V_{B}/V) with one fitted constant AA was used.

Finally, to calculate the number of excitons (and thus the blocked volume), one can consider the power to sustain a single exciton

P1=EjτjP_{1}=\frac{E_{j}}{\tau_{j}} (43)

where EjE_{j} and τj\tau_{j} are the energy and lifetime of excitonic state. The number of excitons NeN_{e} is

Ne=PAP1N_{e}=\frac{P_{A}}{P_{1}} (44)

where PAP_{A} is the absorbed laser power; for a sufficiently thick sample, it is equal to the total laser power.

As a first verification of the presented theoretical description, one can examine the results obtained in the asymptotic limit of a very large thickness, e.g. a bulk crystal. The results of such a comparison are presented on the Fig.2. Specifically, Fig. 2 a) depicts the calculated optical density spectrum in the region of j=611j=6-11 excitonic states, for two illumination powers. One can notice a quick decrease of absorption in the high power regime, approximately proportional to blockade volume j7\sim j^{7}. This is the so-called optical bleaching [1]. The same result can be seen on the Fig. 2 b), where calculations are compared to the experimental data from [1]. A very good agreement obtained in a wide range of powers and across multiple excitonic states indicates that the saturation model in Eq. (42) is sufficiently precise.

a)Refer to caption b)Refer to caption

Figure 2: a) Calculated optical density, compared to experimental results in bulk crystal [10]; b) Absorption spectrum compared to the data from [1].

V Self-Kerr nonlinearity

In the self Kerr effect the refractive index is changed due to the response of the incoming field itself, in other words it consists in the change of the refractive index of the medium with a variation of the propagating light intensity. The third-order nonlinear susceptibility is the basis of theoretical description of this phenomenon. The nonlinear optical response is conveniently described in terms of a field-dependent index n(E)n(E) defined as

n2(E)=1+χ=ϵb+χ(1)+χ(3)|E|2+.n^{2}(E)=1+\chi=\epsilon_{b}+\chi^{(1)}+\chi^{(3)}|E|^{2}+.... (45)

The real part of the nonlinear susceptibility defines the nonlinear index of refraction, which characterizes so-called Kerr media, n2=Reχ(3)cϵ0n02n_{2}=\frac{Re\chi^{(3)}}{c\epsilon_{0}n_{0}^{2}}, with n02=1+χ(1)n_{0}^{2}=1+\chi^{(1)}.
The self-Kerr interaction is an optical nonlinearity that produces a phase shift proportional to the square of the field intensity (or a number of photons in the field). In the Kerr medium the phase of an electromagnetic wave propagating at the distance LL increases and the increment in phase due to intensity-dependent term is proportional to the distance and to the square of the electric field strength, which is called self-phase modulation. The phase shift is calculated from

ΔΦ=ωLc[n(|E|2)n(0)]\Delta\Phi=\frac{\omega L}{c}[n(|E|^{2})-n(0)] (46)

The considerable nonlinear susceptibility of Rydberg excitonic system, further amplified in a thin quantum well, is expected to cause a noticeable phase shift even for small L100L\sim 100 nm. The confinement states, even when not directly visible, still contribute to the total height of the excitonic line, increasing χ(3)\chi^{(3)} and phase shift.

VI Results

Due to the limited amount of experimental data regarding nonlinear properties of Cu2O quantum wells, as a first step we verify our calculations with a comparison to a bulk medium. It should be stressed that while the calculated spectra approach the bulk ones as LL\rightarrow\infty, the presented method is derived under the assumptions of strong confinement and long-wave approximation, so it yields fully correct results only for well thickness significantly below 1μ1\mum. In contrast to [14], where weak confinement regime is studied, the confinement affects the relative motion of the electron-hole pair.

As mentioned above, one can make a rough comparison with experimental results in bulk medium by assuming a large value of LL, skipping the wide quantum well regime at moderate L1L\sim 1 μ\mum. The linear and nonlinear parts of susceptibility have been calculated from Eqs. (II.2) and (37), in a wide range of laser powers and for a thick crystal L=50L=50 μ\mum. The results are shown on the Fig. 3.

a)Refer to caption b)Refer to caption

Figure 3: Imaginary parts of linear χ(1)\chi^{(1)} and nonlinear χ(3)\chi^{(3)} susceptibility in a bulk crystal (L=50μL=50~{}\mum), for selected laser powers.

The calculated spectra are in the range from j=3j=3 exciton resonance (2.161 meV) to the band gap (2.172 meV). The effect of Rydberg blockade is included in calculations by multiplying the obtained susceptibility by the factor pep_{e}, Eq.(42); as the power increases, the density of excitons reaches saturation and pe0p_{e}\rightarrow 0. In such a way one is able to control whether one is still in the regime, in which additional effects due to Rydberg blockade preventing the transmission are absent, and do not influence the excitons-light interaction. One can see that the overall amplitude of the linear susceptibility changes in the order of 10410^{-4} and the nonlinear part is approximately 3 orders of magnitude lower. As expected, the number of observed resonances is strongly dependent on the power P=12cSbϵ|E|2P=\frac{1}{2}cS_{b}\epsilon|E|^{2}, where Sb=0.1S_{b}=0.1 mm2 is the beam area; for P=1P=1 W, the bleaching is considerable even for j=3j=3. The results are consistent with our previous calculations in a bulk medium [10, 8], as well as experimental observations [30] and indicate that the model in Eq. (42) is correct.

As the next step, let us consider a thin L=100L=100 nm quantum well. In such a system, one can expect that the absorption spectrum will contain multiple confinement states corresponding to the quantum number N=1,2,3N=1,2,3.... This is the case shown in Fig. 4. Although there is no strict upper limit on the confinement state number NmaxN_{max}, in practice only a few lowest confinement states are observable and thus in calculations one can assume Nmax=10N_{max}=10. The linear part of susceptibility is consistent with the results presented in [15] for the case of a quantum well. Specifically, one can see a series of secondary peaks originating from every excitonic line, which shift towards high energy as LL becomes very small. It should be stressed that these lines corresponding to confinement states are only detectable in the case of a very thin quantum well; in the micrometer-sized nanoparticles, the energy spacing between these lines is small enough that they completely overlap, forming a single, broadened excitonic line [17]. Moreover, in this size range, one cannot observe oscillations of the absorption coefficient caused by the spatial matching between the center-of-mass exciton motion and light waves [18]. Naturally, a very small absorption of a thin sample makes a direct observation of confinement states challenging. Moreover, just like in the case of large quantum dots [17], the oscillator strength of excitonic states decreases faster than j3j^{-3}; this is also consistent with the observations in [14]. This effect, in addition to the broadening and chaotic ,,background” formed by multiple confinement lines, puts an upper limit on the maximum principal number of the observable state.

a)Refer to caption b)Refer to caption

Figure 4: Imaginary parts of linear χ(1)\chi^{(1)} and nonlinear χ(3)\chi^{(3)} susceptibility in L=100L=100 nm quantum well, for selected laser powers.

The nonlinear susceptibility shown in the Fig. 4 b) is apparently similar to the bulk case in the Fig. 3. The influence of the confinement on the nonlinear part χ(3)\chi^{(3)} is complex. One can see from Eq. (III) that oscillating terms of VNe,NhV_{N_{e},N_{h}} interplay with slowly varying factors λthe,h\lambda_{the,h}, describing plasma effects, resulting in an absorption attenuation. Namely, the confinement lines are much less pronounced and only N=1N=1 line is readily visible. This effect follows from Eq. (III).

For low-dimensional systems the nonlinear optical effects depend strongly on the shape of the confinement potentials. For the above-used no-escape boundary conditions we obtained the expressions VNe,NhV_{N_{e},N_{h}} decaying as N1N^{-1}. The physics behind is that the rapid motion of electrons and holes in the confinement in zz direction, especially for states with higher NN, hinders the creation of plasma which is responsible for the reduction of the absorption while the linear absorption does not depend on NN. For low-dimensional systems the nonlinear optical effects depend strongly on the shape of the confinement potentials. In the considered quantum well with no-escape boundary conditions, the overall amplitude of the nonlinear part of the susceptibility is enhanced as compared to bulk system. The influence of the confinement on the nonlinear part of χ(3)\chi^{(3)} is more complex. One can see that oscillating functions VNeNhV_{{N_{e}}{N_{h}}}, characteristic for low dimensional confined systems interplay with relatively slowly varying exponential functions due to plasmonic terms, which results in increasing of the nonlinear absorption.

The next Fig. 5 shows the susceptibility spectra calculated for low a laser power and various values of thickness.

a)Refer to caption b)Refer to caption

Figure 5: Imaginary parts of linear χ(1)\chi^{(1)} and nonlinear χ(3)\chi^{(3)} susceptibility for selected values of quantum well thickness, for P=0.1 mW.

One can see that both confinement lines and the main excitonic lines are blueshifted in the limit of small LL; as noted in [14], the confined exciton gains additional energy and this energy shift is most pronounced for L<4aBL<4a_{B}, which is approximately 100 nm for n=10n=10 exciton. On the other hand, it is known that excitons cannot form in quantum dots when the dot size r<0.4aBr<0.4a_{B} [14] which indicates the lower limit of applicability of our theoretical description. As before, the lines corresponding to the confinement states are mostly invisible in the nonlinear susceptibility spectrum. In the linear part, one can see that peaks due to those states, located closely to those due to main excitonic states at L=100L=100 nm , shift quickly towards higher energy for smaller LL because of the changing proportion between confinement energy and excitonic state energy. Due to this divergence, a considerable mixing of states occurs and also many lines can be visible in the energy region above the band gap. As mentioned before, the nonlinear part of susceptibility is enhanced in a thin quantum well; on the Fig. 5 b) one can observe that absorption peaks become higher as LL decreases. The confinement of electrons and holes in a QW results in, illustratively speaking, ”squeezing” of excitons, which increases the binding energy and the oscillator strength of excitons, thus leading to an enhancement of the absorption.

Finally, we can explore the real part of susceptibility and the associated Kerr shift. Naturally, as follows from the Kramers-Kronig relations, each peak in the absorption spectrum corresponds to a region of anomalous dispersion, where ReRe χ\chi and thus also phase shift changes sign. This is visible in the Fig. 6 a). As mentioned before, the confinement states are barely visible in the nonlinear part of susceptibility and thus the spectrum is dominated by lines corresponding to excitonic states j=0,1,2j=0,1,2.... Again, we see a divergence towards higher energy as LL decreases and also a reduction of phase shift in the limit of small LL due to the reduced optical length nLnL in Eq. 46. Even for a relatively low thickness L<100L<100 nm, one can observe a phase shift on the order of 50 mrad. The dependence of phase shift on the laser power is shown on the Fig. 6 b). Overall, the lower excitonic states provide a larger phase shift due to their larger oscillator strengths. The shift increases with power but is limited by optical bleaching caused by Rydberg blockade; one can see that the influence of higher states vanishes at high power. On the other hand, in the relatively lower power regime, the stronger nonlinear properties of upper states result in a considerable phase shift.

a)Refer to caption b)Refer to caption

Figure 6: The self-Kerr shift as a function of a) quantum well thickness and b) light power.

An useful measure of the nonlinearity is the maximum phase shift that can be obtained throughout the whole spectrum. The results calculated for a range of input powers are shown on the Fig. 7 a).

a)Refer to caption b)Refer to caption

Figure 7: The maximum value of self-Kerr phase shift as a function of a) light power and b) quantum well thickness.

As expected, the power dependence is linear due to the |E|2P|E|^{2}\sim P factor in Eq. (46). However, the dependence on thickness, shown on the Fig. 7 b), is more complicated. Initially, as LL increases, the phase shift is also rapidly increasing, starting from ΔΦ(L0)=0\Delta\Phi(L\rightarrow 0)=0. However, at some point, the increase of the optical length is compensated by the decrease of χ(3)\chi^{(3)}, which is enhanced in very thin wells. Thus, the phase shift reaches a local maximum and then starts decreasing with increasing LL. Eventually, in the region of L300L\sim 300 nm, the value of χ(3)\chi^{(3)} stabilizes on the same level as in bulk medium and the phase shift again becomes linearly dependent on optical length. One can also notice slight oscillations on the Fig. 7 b) in the region L100L\sim 100 nm. In this regime, the confinement states are mostly visible; the thickness-dependent overlapping of multiple states slightly affects the maximum value of the susceptibility and thus the phase shift. In conclusion, the choice of quantum well thickness, input power and specific excitonic state to realize a self-Kerr shift is highly nontrivial, with multiple tradeoffs influenced by amplification of nonlinear properties, overlap of confinement states and Rydberg blockade.

VII Conclusions

In summary, we have studied the nonlinear interaction between an electromagnetic wave and Rydberg excitons in Cu2O quantum well using the Real Density Matrix Approach, incorporating the control of the Rydberg blockade. Our theoretical, analytical results for linear and nonlinear absorption are illustrated by numerical calculations and indicate the potential experimental conditions for the best observation of confinement states in linear and nonlinear optical spectra. We show that a clear separation of confinement states and an amplification of nonlinear properties of the system are possible in sufficiently thin (L<100L<100 nm) quantum wells. The interplay between nonlinearity enhancement and optical length of the system is discussed. We theoretically demonstrate that the Kerr nonlinearity and significant self-phase modulation are accomplished in a semiconductor quantum well with REs.
In short, our work provides insights into the nonlinear interactions of RE with photons in quantum-confined systems, opening interesting opportunities to explore Rydberg excitons for future opto-electronic nanoscale applications. We hope that our results might be useful for future direct integration of Rydberg confined states with nanophotonic devices.

References

  • [1] T. Kazimierczuk, D. Fröhlich, S. Scheel, H. Stolz, and M. Bayer, Nature 514, 344 (2014).
  • [2] J. Heckötter, M. Freitag, D. Fröhlich, M. Aßmann, M. Bayer, M. A. Semina, and M. M. Glazov, Phys. Rev.B 95, 035210 (2017).
  • [3] M. Assmann, and M. Bayer, Adv. Quantum Technol. 3, 1900134 (2020).
  • [4] S. A. Lynch, C. Hodges, S. Mandal, W. Langbein, R. P. Singh, L. Gallagher, J. D. Pritchett, D. Pizzey, J. P. Rogers, C. Adams, and M. P. Jones, Phys. Rev. Materials 5, 084602 (2021).
  • [5] J. Heckötter, V. Walther, S. Scheel, M. Bayer, T. Pohl, and M. Assmann, Nature Communications 12, 3556 (2021).
  • [6] L.A.P. Gallagher, J.P. Rogers, J.D. Pritchett, R.A. Mistry, D. Pizzey, Ch.S. Adams, M.P.A. Jones, P. Grünwald, V. Walther, Ch. Hodges, and W. Langbein, and S. A. Lynch, Phys. Rev. Research 4, 013031 (2022).
  • [7] K. Orfanakis, S. Rajendran, V. Walther, T. Volz, T. Pohl, and H. Ohadi, Nature Materials (2022).
  • [8] S. Zielińska-Raczyńska, G. Czajkowski, K. Karpiński, and D. Ziemkiewicz, Phys. Rev. B 99, 245206 (2019).
  • [9] V. Walther, P. Grünwald, and T. Pohl, Phys. Rev. Lett. 125, 173601 (2020).
  • [10] C. Morin, J. Tignon, J. Mangeney, S. Dhillon, G. Czajkowski, K. Karpiński, S. Zielińska-Raczyńska, and D. Ziemkiewicz, T. Boulier, arXiv:2202.09239v1 [quant-ph].
  • [11] V. Walther, R. Johne, and T. Pohl, Nature Communications 9, 1309 (2018).
  • [12] D. Ziemkiewicz, and S. Zielińska-Raczyńska, Optics Letters 43, 3742-3745 (2018).
  • [13] D. Ziemkiewicz, and S. Zielińska-Raczyńska, Optics Express 27(12), 16983 (2019).
  • [14] A. Konzelmann, B. Frank, and H. Giessen, J. Phys. B 53, 024001 (2020).
  • [15] D. Ziemkiewicz, K. Karpiński, G. Czajkowski, and S. Zielińska-Raczyńska, Phys. Rev. B 101, 205202 (2020).
  • [16] D. Ziemkiewicz, G. Czajkowski, K. Karpiński, and S. Zielińska – Raczyńska, Phys. Rev. B 103, 035305 (2021).
  • [17] K. Orfanakis, S. Rajendran, H. Ohadi, S. Zielińska-Raczyńska, G. Czajkowski, K. Karpiński, and D. Ziemkiewicz, Phys. Rev. B 103, 245426 (2021).
  • [18] M. Takahata, K. Tanaka, and N. Naka, Phys. Rev. B 97, 205305, (2018)
  • [19] S. Steinhauer, M.A.M. Versteegh, S. Gyger, A. Elshaari, B. Kunert, A. Mysyrowicz, and V. Zwiller, Commun Mater 1, 11 (2020). https://doi.org/10.1038/s43246-020-0013-6
  • [20] H. R. Hamedi, and M. R. Mehmannavaz, Physica E: Low-dimensional Systems and Nanostructures 66, 309-316, (2015).
  • [21] S. G. Kosionis, A. F. Terzis, and E. Paspalakis, Journal of Applied Physics 109(8), 084312 (2011).
  • [22] C. Zhu, and G. Huang, Opt Express 19(23), 23364-23376 (2011).
  • [23] H. Qian, Y. Xiao, and Z. Liu, Nat Commun 7, 13153 (2016).
  • [24] A. Stahl and I. Balslev, Electrodynamics of the Semiconductor Band Edge (Springer-Verlag, Berlin-Heidelberg-New York, 1987).
  • [25] G. Czajkowski, F. Bassani, and A. Tredicucci, Polaritonic effects in superlattices, Phys. Rev. B 54, 2035 (1996).
  • [26] S. Zielińska – Raczyńska, G. Czajkowski, and D. Ziemkiewicz, Phys. Rev. B 93, 075206 (2016).
  • [27] M. Abramowitz and I. Stegun, Handbook of Mathematical Functions (Dover Publications, New York, 1965).
  • [28] D. Kang, A. Gross, H. Yang, Y. Morita, K. Choi, K. Yoshioka, and N. Y. Kim, Phys Rev. B 103, 205203 (2021).
  • [29] D. Frank and A. Stahl, Solid State Commun. 52, 861 (1984).
  • [30] J. Heckötter, M. Freitag, D. Fröhlich, M. Assmann, M. Bayer, P. Grünwald, F. Schöne, D. Semkat, H. Stolz, and S. Scheel, Phys. Rev. Lett. 121, 097401 (2018).
  • [31] Zielińska-Raczyńska, D. Ziemkiewicz, and G. Czajkowski, Phys. Rev. B 97, 165205 (2018).
  • [32] H. Stolz, F. Schöne, and D. Semkat, New. J. Phys. 20, 023019, (2018).
  • [33] N. Naka, I. Akimoto, M. Shirai, and Ken-ichi Kan’no, Phys. Rev. B 85, 035209 (2012).
  • [34] S. Zielińska-Raczyńska, G. Czajkowski, and D. Ziemkiewicz, Phys. Rev. B 93, 075206 (2016).

Appendix A Coefficients A,B

Denoting by f~0e\tilde{f}_{0e} and f~0h\tilde{f}_{0h} the modified Boltzmann distributions f0ef_{0e} and f)hf_{)h} for electron and holes respectively, projections of f0e,hf_{0e,h} on the eigenfunctions Ψ1NeNh\Psi_{\ell 1N_{e}N_{h}} are given by the following expressions

ΨNeNh|f~0e(r)=ANeNh,\displaystyle\langle\Psi_{\ell N_{e}N_{h}}|\tilde{f}_{0e}(\textbf{r})\rangle=A_{\ell N_{e}N_{h}},
ΨNeNh|f~0h(r)=BNeNh,\displaystyle\langle\Psi_{\ell N_{e}N_{h}}|\tilde{f}_{0h}(\textbf{r})\rangle=B_{\ell N_{e}N_{h}},

which can be used to calculate the constants

ANeNh=ψj(r,ϕ)|f~0e(r,ϕ)ΨNeNh|f~0e(ze,zh),\displaystyle A_{\ell N_{e}N_{h}}=\langle\psi_{j}(r,\phi)|\tilde{f}^{\parallel}_{0e}(r,\phi)\rangle\langle\Psi_{N_{e}N_{h}}|\tilde{f}_{0e}^{\perp}(z_{e},z_{h})\rangle,
=π21κ1/2𝒜(2κ)βM(,3,4κ1ρ0)(+1)(+2)INeNh(e),\displaystyle=\sqrt{\frac{\pi}{2}}\frac{1}{\kappa_{\ell}^{1/2}}{\mathcal{A}}\,(2\kappa_{\ell})^{\beta}\,M\left(-\ell,3,4\kappa_{\ell 1}\rho_{0}\right)\sqrt{(\ell+1)(\ell+2)}\,I_{N_{e}N_{h}}^{(e)},
BNeNh=ψj(r,ϕ)|f~0h(r,ϕ)ΨNeNh|f~0h(ze,zh)\displaystyle B_{\ell N_{e}N_{h}}=\langle\psi_{j}(r,\phi)|\tilde{f}^{\parallel}_{0h}(r,\phi)\rangle\langle\Psi_{N_{e}N_{h}}|\tilde{f}_{0h}^{\perp}(z_{e},z_{h})\rangle
=π21κ1/2(2κ)γM(,3,4κρ0)(+1)(+2)INeNh(h),\displaystyle=\sqrt{\frac{\pi}{2}}\frac{1}{\kappa_{\ell}^{1/2}}{\mathcal{B}}\,(2\kappa_{\ell})^{\gamma}\,M\left(-\ell,3,4\kappa_{\ell}\rho_{0}\right)\sqrt{(\ell+1)(\ell+2)}\,I_{N_{e}N_{h}}^{(h)},

where the following approximation has been used

ψj(r,ϕ)|f~0e(r,ϕ)\displaystyle\langle\psi_{j}(r,\phi)|\tilde{f}^{\parallel}_{0e}(r,\phi)\rangle
=0ρ𝑑ρR1(ρ)π2ρλ~th eexp[ρ22λ~th e2]\displaystyle=\int\limits_{0}^{\infty}\rho\,d\rho\,R_{\ell 1}(\rho)\sqrt{\frac{\pi}{2}}\frac{\rho}{\tilde{\lambda}_{\hbox{\tiny th e}}}\exp\left[-\frac{\rho^{2}}{2\tilde{\lambda}_{\hbox{\tiny th e}}^{2}}\right]
=0ρ𝑑ρπ2ρλ~th eexp[ρ22λ~th e2]\displaystyle=\int\limits_{0}^{\infty}\rho\,d\rho\,\sqrt{\frac{\pi}{2}}\frac{\rho}{\tilde{\lambda}_{\hbox{\tiny th e}}}\exp\left[-\frac{\rho^{2}}{2\tilde{\lambda}_{\hbox{\tiny th e}}^{2}}\right]
×C(4κρ)e2κρM(,3,4κρ)\displaystyle\times C_{\ell}(4\kappa_{\ell}\rho)e^{-2\kappa_{\ell}\rho}M\left(-\ell,3,4\kappa_{\ell}\rho\right)
4κλ~th eπ2CM(,3,4κρ0)0ρ3𝑑ρe2κρρ22λ~th e2\displaystyle\approx\frac{4\kappa_{\ell}}{\tilde{\lambda}_{\hbox{\tiny th e}}}\sqrt{\frac{\pi}{2}}C_{\ell}M\left(-\ell,3,4\kappa_{\ell}\rho_{0}\right)\int\limits_{0}^{\infty}\rho^{3}\,d\rho\,e^{-2\kappa_{\ell}\rho-\frac{\rho^{2}}{2\tilde{\lambda}_{\hbox{\tiny th e}}^{2}}}
=14κ3λ~theπ2CM(,3,4κρ0)exp[f(x,z)]\displaystyle=\frac{1}{4\kappa_{\ell}^{3}{\tilde{\lambda}_{\hbox{\tiny the}}}}\sqrt{\frac{\pi}{2}}C_{\ell}M\left(-\ell,3,4\kappa_{\ell}\rho_{0}\right)\exp[f(x,z)]
=12κ3/2λ~th eπ2(+1)(+2)M(,3,4κρ0)exp[f(x,z)]\displaystyle=\frac{1}{2\kappa_{\ell}^{3/2}{\tilde{\lambda}_{\hbox{\tiny th e}}}}\sqrt{\frac{\pi}{2}}\sqrt{(\ell+1)(\ell+2)}M\left(-\ell,3,4\kappa_{\ell}\rho_{0}\right)\exp[f(x,z)]
f(x,z)=4lnz(e)+3lnx(e)+ln[σ(e)2π]12x(e)2z(e)x(e),\displaystyle f(x,z)=4\ln z_{\ell}^{(e)}+3\ln x_{\ell}^{(e)}+\ln[\sigma_{\ell}^{(e)}\sqrt{2\pi}]-\frac{1}{2}x_{\ell}^{(e)2}-z_{\ell}^{(e)}x_{\ell}^{(e)},

and the integral is evaluated as follows

0ρ3𝑑ρe2κρρ22λ~th e2\displaystyle\int\limits_{0}^{\infty}\rho^{3}\,d\rho\,e^{-2\kappa_{\ell}\rho-\frac{\rho^{2}}{2\tilde{\lambda}_{\hbox{\tiny th e}}^{2}}}
=1(2κ)4(z(e))4Γ(4)e(z(e))2/4D4(z(e)),\displaystyle=\frac{1}{(2\kappa_{\ell})^{4}}(z^{(e)}_{\ell})^{4}\Gamma(4)e^{(z^{(e)}_{\ell})^{2}/4}D_{-4}(z^{(e)}_{\ell}),
z(e)=2κλ~th e=2λ~th e2+3.\displaystyle z^{(e)}_{\ell}=2\kappa_{\ell}\tilde{\lambda}_{\hbox{\tiny th e}}=\frac{2\tilde{\lambda}_{\hbox{\tiny th e}}}{2\ell+3}.

The D4[z(e)]D_{-4}\left[z^{(e)}_{\ell}\right] is the parabolic cylinder function [27], and

z(e)=2κλ~th e=2λ~th e2+3,λ~th e=λth ea.z^{(e)}_{\ell}=2\kappa_{\ell}\tilde{\lambda}_{\hbox{\tiny th e}}=\frac{2\tilde{\lambda}_{\hbox{\tiny th e}}}{2\ell+3},\quad\tilde{\lambda}_{\hbox{\tiny th e}}=\frac{{\lambda}_{\hbox{\tiny th e}}}{a^{*}}.

The term containing function DνD_{\nu} can be approximated as follows

z(e)4Γ(4)ez(e)2/4D4[z(e)]z(e)4σ(e)2πexp{f[x(e)]},\displaystyle z_{\ell}^{(e)4}\Gamma(4)e^{z_{\ell}^{(e)2}/4}D_{-4}\left[z_{\ell}^{(e)}\right]\approx z_{\ell}^{(e)4}\sigma^{(e)}_{\ell}\sqrt{2\pi}\exp\left\{f\left[x_{\ell}^{(e)}\right]\right\},
x(e)=12+z(e)2z(e)2,\displaystyle x_{\ell}^{(e)}=\frac{\sqrt{12+z_{\ell}^{(e)2}}-z_{\ell}^{(e)}}{2},
σ(e)=(1+3x(e)2)1/2,\displaystyle\sigma^{(e)}_{\ell}=\left(1+\frac{3}{x_{\ell}^{(e)2}}\right)^{-1/2},
f[x(e)]=3lnx(e)12x(e)2z(e)x(e).\displaystyle f\left[x_{\ell}^{(e)}\right]=3\ln x_{\ell}^{(e)}-\frac{1}{2}x_{\ell}^{(e)2}-z_{\ell}^{(e)}x_{\ell}^{(e)}.

Appendix B calculation of χ(3)\chi^{(3)}

The equation (37) can be written in the form

χ(3)(ω)=ϵ0(ϵbΔLT)2a3e4ρ0(1Γ01)\displaystyle\chi^{(3)}(\omega)=-\epsilon_{0}(\epsilon_{b}\Delta_{LT})^{2}a^{*3}e^{-4\rho_{0}}\left(\frac{1}{{\mit\Gamma}_{01}}\right)
×jnenhΓjΨNeNhLΨjnenh(r0)fj(ETjnenhω)2+Γj2\displaystyle\times\sum\limits_{jn_{e}n_{h}}{\mit\Gamma_{j}}\langle\Psi_{N_{e}N_{h}}\rangle_{L}\frac{\Psi_{jn_{e}n_{h}}(r_{0})\sqrt{f_{j}}}{(E_{Tjn_{e}n_{h}}-\hbar\omega)^{2}+{\mit\Gamma_{j}}^{2}}
×NeNhfET1NeNh(ANeNh+BNeNh)ETNeNh2(ω+iΓNeNh)2,\displaystyle\times\sum\limits_{\ell N_{e}N_{h}}\frac{\sqrt{f_{\ell}}E_{T1\ell N_{e}N_{h}}(A_{\ell N_{e}N_{h}}+B_{\ell N_{e}N_{h}})}{E_{T\ell N_{e}N_{h}}^{2}-(\hbar\omega+i{\mit\Gamma}_{\ell N_{e}N_{h}})^{2}},

where

ΨNeNh=ψ(r,ϕ)ψL,Ne(1D)(ze)ψL,Nh(1D)(zh),\displaystyle\Psi_{\ell N_{e}N_{h}}=\psi_{\ell}(r,\phi)\psi_{L,N_{e}}^{(1D)}(z_{e})\psi_{L,N_{h}}^{(1D)}(z_{h}),
ψm(r,ϕ)=Rmeimϕ2π=1aeimϕ2π\displaystyle\psi_{\ell m}(r,\phi)=R_{\ell m}\frac{e^{im\phi}}{\sqrt{2\pi}}=\frac{1}{a^{*}}\frac{e^{im\phi}}{\sqrt{2\pi}}
×e2κmr/a(4κmra)m 4κm3/21(2m)![(+2m)!]1/2[!]1/2\displaystyle\times e^{-2\kappa_{\ell m}r/a^{*}}\left(4\kappa_{\ell m}\frac{r}{a^{*}}\right)^{m}\,4\kappa_{\ell m}^{3/2}\frac{1}{(2m)!}\frac{[(\ell+2m)!]^{1/2}}{[\ell!]^{1/2}}
×M(,2|m|+1,4κmra),\displaystyle\times M\left(-\ell,2|m|+1,4\kappa_{\ell m}\frac{r}{a^{*}}\right),
=Φm(ϕ)Cm(4κmra)me2κmr/aM(,2|m|+1,4κmra),\displaystyle=\Phi_{m}(\phi)C_{\ell m}\left(4\kappa_{\ell m}\frac{r}{a^{*}}\right)^{m}e^{-2\kappa_{\ell m}r/a^{*}}\,M\left(-\ell,2|m|+1,4\kappa_{\ell m}\frac{r}{a^{*}}\right),
κm=11+2(+|m|),\displaystyle\kappa_{\ell m}=\frac{1}{1+2(\ell+|m|)},
Cm=1a4κm3/21(2m)![(+2m)!]1/2[!]1/2,\displaystyle C_{\ell m}=\frac{1}{a^{*}}4\kappa_{\ell m}^{3/2}\frac{1}{(2m)!}\frac{[(\ell+2m)!]^{1/2}}{[\ell!]^{1/2}},
ψL,Ne(1D)(ze)=2Lcos[(2Ne1)πzeL],\displaystyle\psi_{L,N_{e}}^{(1D)}(z_{e})=\sqrt{\frac{2}{L}}\cos\left[(2N_{e}-1)\pi\frac{z_{e}}{L}\right],
ψL,Nh(1D)(zh)=2Lcos[(2Nh1)πzhL],\displaystyle\psi_{L,N_{h}}^{(1D)}(z_{h})=\sqrt{\frac{2}{L}}\cos\left[(2N_{h}-1)\pi\frac{z_{h}}{L}\right],
ANeNh=ΨNeNh|f~0e(r),\displaystyle A_{\ell N_{e}N_{h}}=\langle\Psi_{\ell N_{e}N_{h}}|\tilde{f}_{0e}(\textbf{r})\rangle,
ANeNh=ψ(r,ϕ)|f~0e(r,ϕ)ΨNeNh|f~0e(ze,zh),\displaystyle A_{\ell N_{e}N_{h}}=\langle\psi_{\ell}(r,\phi)|\tilde{f}^{\parallel}_{0e}(r,\phi)\rangle\langle\Psi_{N_{e}N_{h}}|\tilde{f}_{0e}^{\perp}(z_{e},z_{h})\rangle,
BNeNh=ψ(r,ϕ)|f~0h(r,ϕ)ΨNeNh|f~0h(ze,zh),\displaystyle B_{\ell N_{e}N_{h}}=\langle\psi_{\ell}(r,\phi)|\tilde{f}^{\parallel}_{0h}(r,\phi)\rangle\langle\Psi_{N_{e}N_{h}}|\tilde{f}_{0h}^{\perp}(z_{e},z_{h})\rangle,
ΨNeNh=ψL,Ne(1D)(ze)ψL,Nh(1D)(zh),\displaystyle\Psi_{N_{e}N_{h}}=\psi_{L,N_{e}}^{(1D)}(z_{e})\psi_{L,N_{h}}^{(1D)}(z_{h}),
f~0e(r)=f0e~(ρ,ze,zh,ϕ)=π2rλth e\displaystyle\tilde{f}_{0e}(\textbf{r})=\tilde{f_{0e}}(\rho,z_{e},z_{h},\phi)=\sqrt{\frac{\pi}{2}}\frac{r}{\lambda_{\hbox{\tiny th e}}}
×[Φ1(ϕ)+Φ1(ϕ)]exp(r2+(zezh)22mekB𝒯2)\displaystyle\times[\Phi_{1}(\phi)+\Phi_{-1}(\phi)]\exp\left(-\frac{r^{2}+(z_{e}-z_{h})^{2}}{2}\frac{m_{e}k_{B}{\mathcal{T}}}{\hbar^{2}}\right)
=f~0e(ze,zh)f~0e(r,ϕ),\displaystyle=\tilde{f}_{0e}^{\perp}(z_{e},z_{h})\tilde{f}^{\parallel}_{0e}(r,\phi),
f~0e(ze,zh)=exp((zezh)22mekB𝒯2),\displaystyle\tilde{f}_{0e}^{\perp}(z_{e},z_{h})=\exp\left(-\frac{(z_{e}-z_{h})^{2}}{2}\frac{m_{e}k_{B}{\mathcal{T}}}{\hbar^{2}}\right), (47)
f~0e(r,ϕ)=π2rλth eexp(r22mekB𝒯2)[Φ1(ϕ)+Φ1(ϕ)],\displaystyle\tilde{f}^{\parallel}_{0e}(r,\phi)=\sqrt{\frac{\pi}{2}}\frac{r}{\lambda_{\hbox{\tiny th e}}}\exp\left(-\frac{r^{2}}{2}\frac{m_{e}k_{B}{\mathcal{T}}}{\hbar^{2}}\right)[\Phi_{1}(\phi)+\Phi_{-1}(\phi)],
r=x2+y2,\displaystyle r=\sqrt{x^{2}+y^{2}},

and

Φm(ϕ)=eimϕ2π,\displaystyle\Phi_{m}(\phi)=\frac{e^{im\phi}}{\sqrt{2\pi}},
λth e=(2mekB𝒯)1/2=2μmeRkB𝒯a,\displaystyle\lambda_{\hbox{\tiny th e}}=\left(\frac{\hbar^{2}}{m_{e}k_{B}{\mathcal{T}}}\right)^{1/2}=\sqrt{\frac{2\mu}{m_{e}}}\sqrt{\frac{R^{*}}{k_{B}{\mathcal{T}}}}a^{*},

is the so-called thermal length (here for electrons).

Similarly, for the hole equilibrium distribution, we have

f~0h(r)=f0h~(ρ,ze,zh,ϕ)=π2rλth h\displaystyle\tilde{f}_{0h}(\textbf{r})=\tilde{f_{0h}}(\rho,z_{e},z_{h},\phi)=\sqrt{\frac{\pi}{2}}\frac{r}{\lambda_{\hbox{\tiny th h}}}
×[Φ1(ϕ)+Φ1(ϕ)]exp(r2+(zezh)22mhkB𝒯2)=\displaystyle\times[\Phi_{1}(\phi)+\Phi_{-1}(\phi)]\exp\left(-\frac{r^{2}+(z_{e}-z_{h})^{2}}{2}\frac{m_{h}k_{B}{\mathcal{T}}}{\hbar^{2}}\right)=
=f~0h(ze,zh)f~0h(r,ϕ),\displaystyle=\tilde{f}_{0h}^{\perp}(z_{e},z_{h})\tilde{f}^{\parallel}_{0h}(r,\phi), (48)
f~0h(ze,zh)=exp((zezh)22mhkB𝒯2),\displaystyle\tilde{f}_{0h}^{\perp}(z_{e},z_{h})=\exp\left(-\frac{(z_{e}-z_{h})^{2}}{2}\frac{m_{h}k_{B}{\mathcal{T}}}{\hbar^{2}}\right),
f~0h(r,ϕ)=π2rλth hexp(r22mhkB𝒯2)[Φ1(ϕ)+Φ1(ϕ)],\displaystyle\tilde{f}^{\parallel}_{0h}(r,\phi)=\sqrt{\frac{\pi}{2}}\frac{r}{\lambda_{\hbox{\tiny th h}}}\exp\left(-\frac{r^{2}}{2}\frac{m_{h}k_{B}{\mathcal{T}}}{\hbar^{2}}\right)[\Phi_{1}(\phi)+\Phi_{-1}(\phi)],

with the hole thermal length

λth h=(2mhkB𝒯)1/2=2μmhRkB𝒯a,\displaystyle\lambda_{\hbox{\tiny th h}}=\left(\frac{\hbar^{2}}{m_{h}k_{B}{\mathcal{T}}}\right)^{1/2}=\sqrt{\frac{2\mu}{m_{h}}}\sqrt{\frac{R^{*}}{k_{B}{\mathcal{T}}}}a^{*},

with the radial part ψj(r,ϕ)|f~0e(r,ϕ)\psi_{j}(r,\phi)|\tilde{f}^{\parallel}_{0e}(r,\phi)\rangle defined in Appendix A.

Appendix C table of parameters

Table 1: Band parameter values for Cu2O, masses in free electron mass m0m_{0}.
Parameter Value Unit Reference
EgE_{g} 2172.08 meV [1]
RR^{*} 87.78 meV [31]
ΔLT\Delta_{LT} 1.25×1031.25\times 10^{-3} meV [32]
mem_{e} 0.99 m0m_{0} [33]
mhm_{h} 0.58 m0m_{0} [33]
μ\mu 0.363 m0m_{0}
μ\mu^{\prime} -2.33 m0m_{0}
MtotM_{tot} 1.56 m0m_{0}
aa^{*} 1.1 nm [31]
r0r_{0} 0.22 nm [34]
ϵb\epsilon_{b} 7.5 [1]
T1T_{1} 500 ns