1: \documentstyle[aps,prb,preprint]{revtex}
2: \renewcommand{\baselinestretch}{1}
3: \begin{document}
4: \draft
5: \title{Kinetics of exciton photoluminescence in type-II
6: semiconductor superlattices}
7:
8: \author{L.~S.~Braginsky\thanks{Electronic address:
9: brag@isp.nsc.ru}, M.~Yu.~Zaharov, A.~M.~Gilinsky,
10: V.~V.~Preobrazhenskii, M.~A.~Putyato, and K.~S.~Zhuravlev}
11: \address{Institute of Semiconductor Physics, 630090,
12: Novosibirsk, Russia}
13:
14: \date{\today}
15: \maketitle
16: \begin{abstract}
17: The exciton decay rate at a rough
18: interface in type-II semiconductor superlattices is
19: investigated. It is shown that the possibility of recombination
20: of indirect excitons at a plane interface essentially affects
21: kinetics of the exciton photoluminescence at a rough interface.
22: This happens because of strong correlation between the
23: exciton recombination at the plane interface and at the
24: roughness. Expressions that relate the parameters of the
25: luminescence kinetics with statistical characteristics of the
26: rough interface are obtained. The mean height and
27: length of roughnesses in GaAs/AlAs superlattices are estimated
28: from the experimental data.
29: \end{abstract}
30: \pacs{78.66.Fd}
31: \narrowtext
32: \section{Introduction}
33: GaAs/AlAs type-II superlattices are the subject of extensive
34: investigations in the recent decade. Electrons and holes are
35: separated in these structures: holes are confined in the
36: $\Gamma$ valley of GaAs, whereas electrons are in $X$ valleys
37: of AlAs. Changing the width of AlAs layer during the growth
38: of the structure, it is possible to confine the electrons
39: either in the $X_z$ valley ($X$ valley that is directed
40: along the structure axis [001]) or in the $X_{xy}$ valley
41: ($X$ valley that is directed along the GaAs/AlAs interface:
42: [100] or [010]). The excitons in such structures are indirect
43: in both real and momentum spaces.
44:
45: Kinetics of the exciton luminescence is
46: investigated by the time-resolved method.
47: The theory by Klein {\it et al. } \cite{Klein} is usually used
48: to explain the results of such experiments. The theory
49: has been developed to consider the no-phonon
50: radiative decay rates of indirect excitons in alloy
51: semiconductors (e.g., Ga$_{1-x}$Al$_x$As). The recombination of
52: indirect excitons occurs because of intervalley scattering of
53: electrons at the potential fluctuations caused by the
54: compositional disorder. These short-range scatterers are
55: necessary to compensate the large momentum of the electron in
56: the $X$ valley. The nonexponential dependence of the decay rate
57: has been obtained
58: \begin{equation}
59: \label{1}
60: I(t)\propto e^{-w_0t}(1+2w_rt)^{-3/2}.
61: \end{equation}
62: Where the value $w_r$ is connected with the compositional
63: disorder. The exponential factor has been included in
64: Eq.~(\ref{1}) to consider different nonstochastic processes of
65: the exciton recombination (e.g., the phonon-assisted
66: recombination). This is possible only in the absence of
67: correlation between stochastic and nonstochastic processes.
68:
69: The possibility to apply the theory \cite{Klein} for
70: superlattices has been discussed by F.~Minami
71: {\it et al.}.~\cite{Minami} Authors suppose the short-range
72: scatterers are distributed along the plane boundary. This
73: assumption justifies the application of Eq.~(\ref{1}) for
74: superlattices; however, it does not allow to relate the
75: parameter $w_r$ with characteristics of the rough interface,
76: e.g., the mean height and length of roughnesses. I. Krivorotov
77: {\it et al.}\cite{Krivirotov} have showed that nonradiative
78: decay due to exciton trapping by interfacial defects also
79: results in nonexponential factor in Eq.~(\ref{1}).
80: Nevertheless, Eq.~(\ref{1}), wherein the parameters $w_r$ and
81: $w_0$ are considered as trial, is commonly used for explanation
82: of the experimental results.\cite{Nagao}
83:
84: It should be noted that roughnesses are not necessary for
85: the recombination of $X_z$ excitons. Their recombination occurs
86: even at a plain interface where the normal component of the
87: electron momentum relaxes. This important
88: point also distinguishes the exciton recombination in
89: superlattices. The process, however, can not be taken into
90: account by a simple exponential factor. Indeed, the wave function
91: of the electron at a rough interface is the
92: sum of its regular and diffuse components. The first one exists
93: at a plane interface, whereas the latter is due to the
94: roughnesses. For this reason the crossed terms arise in
95: the interband matrix element; so that the probability of the
96: exciton recombination, which is determined by the squared module
97: of this matrix element, is no longer a simple sum of the
98: probabilities of the recombination at the plane
99: interface and at the roughnesses. This correlation leads to a
100: more complicated relation than the simple exponential factor in
101: Eq.~(\ref{1}).
102:
103:
104:
105: In this paper we consider a more realistic model of the rough
106: interface. We show that Eq.~(\ref{1}) holds for
107: the decay rate of $X_{xy}$ excitons and relate
108: the $w_r$ value with parameters of the rough interface. We
109: determine the decay rate of $X_z$ excitons. In particular,
110: it is found that this value at large times behaves roughly as
111: $I(t)\propto \exp (-w_0t)/t$, rather than $I(t)\propto \exp
112: (-w_0t)/t^{3/2}$ as it is predicted by Eq.~(\ref{1}).
113:
114: Our experiments on the GaAs/AlAs type-II superlattices confirm
115: these results. We use the experimental data for the radiative
116: decay rates to estimate the parameters of the rough interface.
117: The mean height of roughnesses was found to be close to
118: the lattice constant, whereas their mean length
119: is about $50\,\AA$.
120:
121: \section{Radiative decay rates of indirect excitons in
122: superlattices. Theory}
123:
124: Let $z=0$ be the interface between GaAs ($-d_1<z<0$) and AlAs
125: ($0<z<d_2$), and $\bbox{\rho}$ be the vector in the $XY$ plane.
126: We consider the exiton recombination at the interface
127: and write the exciton wave function as
128: follows:\cite{Ivchenko}
129: \[
130: \phi({\bf r}_e,{\bf r}_h)=f_e(z_e)f_h(z_h)
131: G(\bbox{\rho}_e-\bbox{\rho}_h,z_e,z_h).
132: \]
133: Where ${\bf r}_e=\{\bbox{\rho}_e,\ z_e\}$ and ${\bf
134: r}_h=\{\bbox{\rho}_h,\ z_h\}$ are coordinates of the electron
135: and the hole, $f_e(z_e)$ and $f_h(z_h)$ are their wave
136: functions in the absence of Coulomb interaction; the
137: function $G$ takes into account this interaction. The
138: probability for the exciton recombination is proportional to
139: $G^2({\bf 0})$
140: [$G({\bf 0})\equiv G(\bbox{\rho}_e=\bbox{\rho}_h,z_e=z_h=0)$]
141: and the squared module of the matrix element
142: \begin{equation}
143: \label{2}
144: {\cal P}= \int
145: f_e(z)\bbox{\nabla}f_h(z)\,dz\,d^2\bbox{\rho}.
146: \end{equation}
147:
148: The functions $f_e(z_e)$ and $f_h(z_h)$ can be expressed via
149: envelope wave functions of the electron and the hole in the
150: conduction and valence bands of GaAs and AlAs. To determine the
151: envelopes, the appropriate boundary conditions at the GaAs/AlAs
152: interface should be imposed. The roughness of the interface
153: has an influence on these boundary conditions and, therefore,
154: affects the envelopes. We shall consider the rough interface
155: where the mean height of roughnesses is small in comparison
156: with the electron wavelength (or the exciton size). This
157: allows us to use the boundary conditions at the rough
158: interface\cite{PhysicaE} to consider an influence of
159: roughnesses on the exciton recombination.
160:
161: \subsection{Boundary conditions for the envelope wave functions
162: at a GaAs/AlAs interface}
163: \subsubsection{Boundary conditions at a plane interface}
164: In general, the boundary conditions for the electron
165: envelopes can be written as follows:
166: \begin{equation}
167: \label{3}
168: \left(
169: \begin{array}{l}
170: \Psi_\Gamma^r\\
171: \Psi_\Gamma^{r'}\\
172: \Psi_{Xxy}^r\\
173: \Psi_{Xxy}^{r'}\\
174: \Psi_{Xz}^r\\
175: \Psi_{Xz}^{r'}
176: \end{array}
177: \right)
178: =\tilde{T}
179: \left(
180: \begin{array}{l}
181: \Psi_\Gamma^l\\
182: \Psi_\Gamma^{l'}\\
183: \Psi_{Xxy}^l\\
184: \Psi_{Xxy}^{l'}\\
185: \Psi_{Xz}^l\\
186: \Psi_{Xz}^{l'}
187: \end{array}
188: \right).
189: \end{equation}
190: Where $\Psi_{\Gamma,Xxy,Xz}^{l,r}$ are the envelopes which
191: correspond to the $\Gamma$, $X_{xy}$, and $X_z$ valleys of GaAs
192: and AlAs; $\Psi_{\Gamma,Xxy,Xz}^{l,r'}\equiv \partial
193: \Psi_{\Gamma,Xxy,Xz}^{l,r}/\partial z$ are their normal
194: derivatives. The elements $\tilde{t}_{ik}$ of the $6\times 6$
195: matrix $\tilde{T}$ are determined by the interface structure.
196: They are independent of the electron energy. For the GaAs/AlGaAs
197: interface they have been calculated by Ando {\it et
198: al.}.\cite{Ando}
199:
200: We shall consider the particular cases of $X_z$ and $X_{xy}$
201: excitons. This allows us to simplify Eq.~(\ref{3}).
202: First, we omit mixing between $X_z$ and $X_{xy}$ valleys.
203: Second, the energy position of $\Gamma$ minimum in AlAs is
204: considerably higher than that of $X$ minimum. For this reason
205: the wave function $\Psi_\Gamma^r$ decays rapidly apart of the
206: interface. We have $\Psi_\Gamma^r\propto \exp(-\gamma^r z),\
207: \Psi_\Gamma^{r'}=-\gamma^r\Psi_\Gamma^{r}$, where
208: $\gamma^r=\sqrt{2m_{\Gamma}^r(E_\Gamma-\varepsilon_e)}$ (here
209: $m_{\Gamma}^r$ is effective mass in the $\Gamma$ valley of
210: AlAs, $\varepsilon_e\approx E_X$ is the electron energy,
211: $E_\Gamma$ and $E_X$ are energies of bottoms of the $\Gamma$ and
212: $X$ valleys) can be considered as independent of the electron
213: energy. By eliminating of $\Psi_\Gamma^{r}$ from Eq.~(\ref{3}),
214: for $X_z$ electrons we find:
215: \begin{mathletters}
216: \label{bcplane}
217: \begin{equation}
218: \label{4}
219: \left\{
220: \begin{array}{l}
221: \Psi_{Xz}^r=\Psi_{Xz}^l,\\
222: \Psi_{Xz}^{r'}=
223: t^z_{41}\Psi_\Gamma^{l}+t^z_{44}\Psi_{Xz}^{l'},\\
224: \Psi_\Gamma^l+t^z_{12}\Psi_\Gamma^{l'}+ t^z_{13}\Psi_{Xz}^l=0.
225: \end{array}
226: \right.
227: \end{equation}
228: Where $t^z_{44}\approx m_{Xl}^r/m_{Xl}^l\approx 1$, this value
229: takes into account the difference of longitudinal effective
230: masses in the $X$ valleys of AlAs and GaAs; $t^z_{41}=t_{\Gamma
231: X} m^r_{Xl}/(m_ea)$, $t^z_{12}=m^r_\Gamma/(m^l_\Gamma\gamma^r)$,
232: $t^z_{13}=t_{\Gamma X}m^l_\Gamma/(m_ea\gamma^r)\ll 1$;
233: $t_{\Gamma X}\approx 1$ is the parameter of $\Gamma$--$X$
234: mixing. Other elements of the ${t}^z_{ik}$ matrix are small;
235: this is the result of numerical calculations.\cite{Ando}
236:
237: Note that the band states in the $X$ valley result from
238: interaction of two close-lying bands: lower $X_1$ and upper
239: $X_3$; meanwhile only the $X_3$ states mix effectively with
240: $\Gamma$ states. This means that $t_{\Gamma X}\approx 1$ is
241: the upper estimation of $\Gamma$--$X$ mixing.
242:
243: It is sufficient to consider only $X$ valleys of each
244: contacted material when $X_{xy}$ electrons are investigated.
245: Assuming $\Psi_\Gamma^{l'}=\gamma^l\Psi_\Gamma^{l}$, where
246: $\gamma^l\sim 2\pi/a$ ($a$ is the lattice constant), from
247: Eq.~(\ref{3}) we find:
248: \begin{eqnarray}
249: \label{5}
250: &&\Psi_{Xxy}^r=t^{xy}_{11}\Psi_{Xxy}^l+t^{xy}_{12}\Psi_{Xxy}^{l'},\\
251: &&\Psi_{Xxy}^{r'}=t^{xy}_{21}\Psi_{Xxy}^l+t^{xy}_{22}\Psi_{Xxy}^{l'}.
252: \nonumber
253: \end{eqnarray}
254: Where $|t^{xy}_{12}|\ll 1,\ |t^{xy}_{21}|\ll a^{-1},\
255: |t^{xy}_{11}|\approx 1$, and $t^{xy}_{22}\approx
256: m_{Xt}^r/m_{Xt}^l\approx 1$; $m_{Xt}^r$ and $m_{Xt}^l$ are the
257: transversal effective masses of AlAs and GaAs.
258:
259: The bands of the light and heavy holes are splitted due
260: to the size quantization.
261: This allows us to consider only the heavy holes in each material
262: and write the boundary conditions for them as follows:
263: \begin{eqnarray}
264: \label{6}
265: &&\Psi_h^r=t^h_{11}\Psi_h^l+t^h_{12}\Psi_h^{l'},\\
266: &&\Psi_h^{r'}=t^h_{21}\Psi_h^l+t^h_{22}\Psi_h^{l'}.
267: \nonumber
268: \end{eqnarray}
269: Where $\Psi_h^{l,r}$ are the envelopes for the heavy holes in
270: each material.
271: \end{mathletters}
272:
273: \subsubsection{Boundary conditions at a rough interface}
274: We shall consider the model of a rough interface that is
275: presented on Fig.~1. This model is in agreement with optical
276: \cite{Lurssen} and structural \cite{Bechstedt} investigations
277: of GaAs/AlAs interface. The interface looks like an array of
278: the plane areas of the same crystallographic orientation. The
279: random function $z=\xi(\bbox{\rho})$ of the coordinates in the
280: $XY$ plane determines the positions of these areas relative to
281: $z=0$.
282:
283: We assume the average
284: height of roughnesses $h$ to be small in comparison with
285: the electron wavelength. Then it is possible to describe the
286: rough interface by means of the correlation function $W(\bbox
287: {\rho '}, \bbox{\rho ''})=\overline{\xi(\bbox{\rho '})\xi(\bbox{\rho
288: ''})}$. For the homogeneous rough interface $W(\bbox{\rho
289: '}, \bbox{\rho ''})=W(\bbox{\rho '}-\bbox{\rho ''})$, i.e.,
290: the correlation function is the function of one variable:
291: $\bbox{\rho}=\bbox{\rho '}-\bbox{\rho ''}$. There are two
292: parameters that are most important when the statistical
293: properties of a rough interface are considered: $h^2=W(0)$ and
294: the correlation length $l$ --- the mean attenuation length of
295: the correlation function. In our model the correlation length
296: $l$ can be associated with the mean size of the plane area.
297:
298: The special form of the rough interface (Fig.~1) allows us to
299: apply the boundary conditions (\ref{bcplane}), which are
300: applicable at a plane interface, at each plane $z=\xi$.
301: The inequality $|\xi\Psi'|\sim h/\lambda\ll 1$ ($\lambda$ is the
302: electron wavelength) allows to rewrite these boundary conditions
303: at a plane $z=0$. After some algebra we obtain:
304: \begin{mathletters}
305: \label{bcrough}
306: \begin{eqnarray}
307: \label{7}
308: &&\Psi_{Xz}^r=-t^z_{41}\eta(\xi)\xi(\bbox{\rho})\Psi_\Gamma^l
309: +\Psi_{Xz}^l+
310: (1-t^z_{44})\xi(\bbox{\rho})\Psi_{Xz}^{l'},\nonumber\\
311: &&\Psi_{Xz}^{r'}=t^z_{41}\eta(\xi)\Psi_\Gamma^l
312: +t^z_{41}\eta(\xi)\xi(\bbox{\rho})\Psi_\Gamma^{l'}
313: t^z_{44}\Psi_{Xz}^{l'},\\
314: &&\Psi_\Gamma^l+[t^z_{12}+\xi(\bbox{\rho})]\Psi_\Gamma^{l'}
315: +t^z_{13}\eta^*(\xi)\Psi_{Xz}^l
316: +t^z_{13}\eta^*(\xi)\xi(\bbox{\rho})\Psi_{Xz}^{l'},
317: \nonumber
318: \end{eqnarray}
319: for electrons in the $X_z$ valley;
320: \begin{eqnarray}
321: \label{8}
322: &&\Psi_{Xxy}^r=\Psi_{Xxy}^l+(1-t^{xy}_{22})\xi(\bbox{\rho})\Psi_{Xxy}^{l'},\\
323: &&\Psi_{Xxy}^{r'}=t^{xy}_{21}\Psi_{Xxy}+t^{xy}_{22}\Psi_{Xxy}^{l'},
324: \nonumber
325: \end{eqnarray}
326: for electrons in the $X_{xy}$ valley; and
327: \begin{eqnarray}
328: \label{9}
329: &&\Psi_h^r=\Psi_h^l+(1-t^h_{22})\xi(\bbox{\rho})\Psi_h^{l'},\\
330: &&\Psi_h^{r'}=t^h_{21}\Psi_h+t^h_{22}\Psi_h^{l'},
331: \nonumber
332: \end{eqnarray}
333: for the holes. Factor $\eta(\xi)=\exp(2\pi i \xi/a)$ in
334: Eq.~(\ref{7}) takes the two values $\pm 1$ for $\xi=a$ or
335: $\xi=a/2$. It has been introduced in Ref.~\cite{Ivchenko1} to
336: take into account the symmetry properies of the Bloch functions
337: with respect to translation by a single monomolecular layer
338: ($a/2$) along the $z$ axis. The Bloch function of the electron
339: in the $X_z$ valley changes its sigh under this translation
340: whereas the Bloch function of the electron in the $\Gamma$
341: valley does not. Therefore, the parameter $t_{\Gamma X}$ of
342: $\Gamma$--$X$ mixing also should change sigh under such
343: translation. This is not important at a plane interface, but it
344: must be taken into account when the relative positions of some
345: interfaces are considered. We assume
346: $|t^{xy,h}_{21}|\ll a^{-1}$: this is the result of numerical
347: calculations.\cite{Ando}
348: \end{mathletters}
349:
350: Unlike Eqs.~(\ref{bcplane}) the boundary conditions
351: (\ref{bcrough}) contain the terms depended on $\xi$. They would
352: not be important, if $\xi={\rm const}$. Then they relevant to
353: the phase shift of the wave functions due to the shift of the
354: interface. However, these terms become important when $\xi$
355: depends on $\bbox{\rho}$. Interference of the electrons
356: scattered from the neighboring planes in the vicinity of
357: steps (like point 1 on Fig.~1) results in the diffuse component
358: of their wave function. The mean size of the region at the step
359: where the interference occurs is the parallel-to-interface
360: component of the electron wavelength. Hence the ratio of this
361: size to the size of the plane area $l$ characterizes the
362: roughnesses influence on the electrons.
363:
364: We separate the diffuse components
365: $\varphi_{\Gamma,X_z,X_{xy}}^{l,r}$ of the envelope wave
366: functions and write the envelopes as follows:
367: \cite{Bass}
368: \begin{eqnarray}
369: \label{10}
370: &&\Psi_{\Gamma,X_z,X_{xy}}^{l,r}=\Phi_{\Gamma,X_z,X_{xy}}^{l,r}
371: +\varphi_{\Gamma,X_z,X_{xy}}^{l,r}, \\
372: &&\mbox{where}
373: \quad \overline{\varphi_{\Gamma,X_z,X_{xy}}^{l,r}}=0.
374: \nonumber
375: \end{eqnarray}
376:
377: \widetext
378: Using the boundary conditions~(\ref{bcrough}),
379: for the envelopes $\Phi_{\Gamma,X_z,X_{xy}}^{l,r}$ and
380: $\varphi_{\Gamma,X_z,X_{xy}}^{l,r}$ (see Ref.~\cite{PhysicaE}
381: for the details) we obtain
382: \begin{eqnarray*}
383: %\label{11}
384: \Phi_\Gamma^l({\bf r})=T_\Gamma e^{-ip_\Gamma z},\quad
385: &\Phi_{X_z,X_{xy}}^l({\bf r})=
386: T_{X_z,X_{xy}}e^{\gamma_{X_z,X_{xy}}z},\quad
387: &\Phi_{X_z,X_{xy}}^r({\bf
388: r})=e^{-iqz}+R_{X_z,X_{xy}}e^{iqz},%\nonumber\\
389: \end{eqnarray*}
390: \begin{eqnarray*}
391: \Phi_h^l({\bf r})=e^{ipz}+R_he^{-ipz},\quad%\nonumber\\
392: &&\Phi_h^r({\bf r})=
393: T_he^{-\gamma_hz},%\\
394: \end{eqnarray*}
395: \begin{eqnarray}
396: \label{11}
397: \varphi_{\Gamma}^l({\bf r})=
398: \frac{2q}{(2\pi)^2}
399: \int_{-\infty}^{\infty} A_{\Gamma}^l(\bbox{k_\parallel})
400: \tilde{\xi}(\bbox{k_\parallel})
401: e^{i(\bbox{\scriptstyle k_\parallel
402: \rho}-k_\Gamma z)}\,
403: d\bbox{k_\parallel},\nonumber \\
404: %
405: \varphi_{X_z,X_{xy}}^l({\bf r})=
406: \frac{2q}{(2\pi)^2}
407: \int_{-\infty}^{\infty} A_{X_z,X_{xy}}^l(\bbox{k_\parallel})
408: \tilde{\xi}(\bbox{k_\parallel})
409: e^{i\bbox{\scriptstyle k_\parallel
410: \rho}+\ae_{\Gamma,X_z,X_{xy}}z}\, d\bbox{k_\parallel},\nonumber
411: \\
412: \varphi_{X_z,X_{xy}}^r({\bf r})= \frac{2q}{(2\pi)^2}
413: \int_{-\infty}^{\infty} A_{\Gamma,X_z,X_{xy}}^r(\bbox{k_\parallel})
414: \tilde{\xi}(\bbox{k_\parallel})
415: e^{i(\bbox{\scriptstyle k_\parallel
416: \rho}+k_{X_z,X_{xy}}z)}\,
417: d\bbox{k_\parallel}, \nonumber\\
418: \varphi_h^l({\bf r})=
419: \frac{2p}{(2\pi)^2}
420: \int_{-\infty}^{\infty} A_h^l(\bbox{k_\parallel})
421: \tilde{\xi}(\bbox{k_\parallel})
422: e^{i(\bbox{\scriptstyle k_\parallel
423: \rho}-k_hz)}\,
424: d\bbox{k_\parallel}, \nonumber\\
425: \varphi_h^r({\bf r})=
426: \frac{2p}{(2\pi)^2}
427: \int_{-\infty}^{\infty} A_h^r(\bbox{k_\parallel})
428: \tilde{\xi}(\bbox{k_\parallel})
429: e^{i\bbox{\scriptstyle k_\parallel
430: \rho}-\ae_hz}\,
431: d\bbox{k_\parallel}. %\nonumber
432: \end{eqnarray}
433: Where
434: \begin{eqnarray*}
435: %\label{12}
436: &&k_\Gamma(\bbox{k_\parallel})=
437: \sqrt{2m_\Gamma(\varepsilon_e-E_\Gamma^l)-\bbox{k_\parallel}^2},
438: \quad
439: \ae_{X_z}(\bbox{k_\parallel})=
440: \sqrt{2m_{Xl}^l(E_{X_z}^l-\varepsilon_e)+\bbox{k_\parallel}^2},\\
441: &&\ae_{X_{xy}}(\bbox{k_\parallel})=
442: \sqrt{2m_{Xt}^l(E_{X_{xy}}^l-\varepsilon_e)+\bbox{k_\parallel}^2},
443: \quad
444: k_{X_z}(\bbox{k_\parallel})=
445: \sqrt{2m_{Xl}^r(\varepsilon_e-E_{X_z}^r)-\bbox{k_\parallel}^2},\\
446: &&k_{X_{xy}}(\bbox{k_\parallel})=
447: \sqrt{2m_{Xt}^r(\varepsilon_e-E_{X_{xy}}^r)-\bbox{k_\parallel}^2},
448: \quad
449: k_h(\bbox{k_\parallel})=
450: \sqrt{2m_{h}^l(E_h^l-\varepsilon_h)-\bbox{k_\parallel}^2},\\
451: &&\ae_h(\bbox{k_\parallel})=
452: \sqrt{2m_{h}^r(\varepsilon_h-E_h^r)+\bbox{k_\parallel}^2},
453: \quad
454: {\rm Im}\,k_{X_z,X_{xy},h}\geq 0,\\
455: %
456: &&T_\Gamma=\frac{2iqt^z_{13}\eta(\xi)}{t^z_{44}\gamma_{X_z}},\;\;\;
457: T_{X_z}=-\frac{2iq}{t^z_{44}\gamma_{X_z}},\;\;\;
458: R_{X_z}=-1-\frac{2iq}{t^z_{44}\gamma_{X_z}},
459: \nonumber\\
460: &&T_{X_{xy}}=\frac{2iq}{t^{xy}_{21}+t^{xy}_{22}\gamma_{X_{xy}}},\;\;\;
461: R_{X_{xy}}=-1+\frac{2iq}{t^{xy}_{21}+t^{xy}_{22}\gamma_{X_{xy}}},
462: \\
463: &&T_h=-\frac{2ip}{-t^h_{21}+t^h_{22}\gamma_h},\;\;\;
464: R_h=-1-\frac{2ip}{-t^h_{21}+t^h_{22}\gamma_h},
465: \nonumber\\
466: &&A_\Gamma^l=\frac{it^z_{13}\eta(\xi)}{t^z_{44}},\quad
467: %\nonumber\\
468: A_{X_z}^r=-\frac{i}{t^z_{44}}
469: \left(\frac{t_{13}^zt_{41}^z}{\gamma_{X_z}}+
470: 1-t^z_{44}\right),\quad
471: %\nonumber \\
472: A_{X_z}^l=-\frac{k_{X_z}}{\ae_{X_z}{t^{z}_{44}}^2}
473: \left(\frac{t_{13}^zt_{41}^z}{\gamma_{X_z}}+
474: 1-t^z_{44}\right),
475: \nonumber \\
476: &&A_{X_{xy}}^r=i
477: \frac{t^{xy}_{22}\ae_l\ae_{X_{xy}}(1-t^{xy}_{22})}
478: {(t^{xy}_{21}+\ae_{X_{xy}}t^{xy}_{22})
479: (t^{xy}_{22}\ae_l+ik_{X_{xy}})},\qquad
480: %\nonumber\\
481: A_{X_{xy}}^l=k_z
482: \frac{\ae_{X_{xy}}(1-t^{xy}_{22})}
483: {(t^{xy}_{21}+\ae_{X_{xy}}t^{xy}_{22})
484: (t^{xy}_{22}\ae_l+ik_{X_{xy}})},
485: \nonumber\\
486: &&A_h^r=
487: -\frac{k_h(1-t^h_{22})}
488: {t^h_{22}(\ae_ht^h_{22}-t^h_{21})},\qquad
489: %\nonumber\\
490: A_h^l=
491: i\frac{\ae_h(1-t^h_{22})}
492: {\ae_ht^h_{22}-t^h_{21}}.
493: \nonumber
494: \end{eqnarray*}
495: \narrowtext
496: Here
497: $\tilde{\xi}(\bbox{k_\parallel})=\int\xi(\bbox{\rho})
498: e^{-i\bbox{k_\parallel \rho}}\,d\bbox{\rho}$,
499: $\gamma_{X_z,X_{xy},h}=\ae_{X_z,X_{xy},h}(0)$,
500: $p_\Gamma=k_\Gamma(0)$; $E_\Gamma^l$, $E_{X_z}^l$,
501: $E_{X_{xy}}^l$, $E_{X_z}^r$, $E_{X_{xy}}^r$, $E_h^l$, and
502: $E_h^r$ are energies of extrema of the appropriate bands.
503: Integration in Eq.~(\ref{11}) is carried out over the whole
504: plane because $\xi(\bbox{\rho})$ is not periodical function of
505: $\bbox{\rho}$. The values of normal-to-interface components of
506: the wave vectors of the electrons $q$ and holes $p$ are
507: determined by the boundary conditions at the interfaces:
508: $z=-d_1$ for $p$ and $z=d_2$ for $q$ (where $d_1$ and $d_2$ are
509: widths of GaAs and AlAs layers). In general, they depends on
510: the valley under consideration: $\tan
511: \frac{qd_2}{2}=-\frac{q}{\gamma_{X_z}}$ for electrons in the
512: $X_z$ valley, $\tan
513: qd_2=-\frac{2q}{t^{xy}_{22}\gamma_{X_{xy}}+t^{xy}_{21}}$ for
514: electrons in the $X_{xy}$ valley, and $\tan
515: pd_1=-\frac{2p}{t^h_{22}\gamma_h-t^h_{21}}$ for the holes. We
516: assume, however, the strong confinement of electrons and holes
517: in the appropriate layers $\gamma_{X_z,X_{xy},h} \gg p,q$, so
518: that $p\approx \pi/d_1$ and $q\approx \pi/d_2$.
519:
520: The wave function of the electron in the GaAs $\Gamma$ valley is
521: small, $p_\Gamma\ll p$; nevertheless, it is real. This
522: distinguish the short-period GaAs/AlAs superlattices from other
523: type-II structures, where the electron wave function decays
524: rapidly from the interface. The electron density is large in
525: AlAs and small, but almost constant, in GaAs. This small part of
526: the electron density could be essential for the exiton
527: recombination would the effective parameter of $\Gamma$--$X$
528: mixing $t_{13}^z$ be sufficently large.
529:
530:
531: \subsection{Radiative decay rates of indirect excitons at a rough
532: interface}
533:
534: To determine the wave functions $f_e(z)$ and $f_h(z)$, we have
535: to insert the corresponding Bloch amplitudes into expressions
536: for the envelopes $\Psi_e$ and $\Psi_h$ (\ref{10}). For
537: instance, for the $X_z$ exciton at a plane interface, we have
538:
539: \begin{eqnarray}
540: \label{13}
541: &&f_e({\bf r})=\frac{1}{\sqrt{N_1}}\left\{
542: \begin{array}{ll}
543: T_\Gamma u_\Gamma({\bf r})e^{\gamma_\Gamma z}\\
544: +T_{X_z}u_{X}({\bf
545: r})e^{\left(\gamma_{X_z}-\frac{2\pi i}{a}\right)z},& z<0,\\ \\
546: u_{X}^*({\bf r})e^{i\left(q-\frac{2\pi i}{a}\right)z} \\
547: +R_{X_z}u_{X}({\bf r})e^{-i\left(q-\frac{2\pi i}{a}\right)z},&
548: z>0,\\
549: \end{array}
550: \right.\nonumber\\
551: &&\ \\
552: &&f_h({\bf
553: r})=\frac{1}{\sqrt{N_2}}\left\{
554: \begin{array}{ll} v({\bf r})e^{ipz}+R_hv^*({\bf
555: r})e^{-ipz},& z<0,\\ T_hv({\bf r})e^{-\gamma_hz}, &z>0.
556: \end{array}
557: \right. \nonumber
558: \end{eqnarray}
559: Where $N_1$ and $N_2$ are numbers of atoms in the AlAs
560: and GaAs layers, $u_{\Gamma}({\bf r})$, $u_{X}({\bf r})$ and
561: $v({\bf r})$ are Bloch amplitudes of electrons in the $\Gamma$
562: and $X$ valleys, and the holes; we assume these amplitudes to be
563: periodical functions of ${\bf r}$.
564:
565: At a rough interface we have to add also the diffuse components
566: of the wave functions. To do that, we have to multiply
567: $\varphi({\bf r})$ (\ref{11}) by the corresponding Bloch
568: amplitudes. Usually the mean size of Bloch amplitudes is
569: small in comparison with the lattice constant. This allows us
570: to assume that $\bbox{\nabla}$-operator in Eq.~(\ref{2}) acts
571: only on these amplitudes and to separate the integration of them
572: from integration of the envelopes. Then the matrix element
573: (\ref{2}) can be written as ${\cal P}={\cal P}_1+{\cal
574: P}_2+{\cal P}_3$, where
575:
576: \begin{eqnarray}
577: \label{14}
578: &&{\cal P}_1=\sum_{\Gamma,X}U_{\Gamma,X}\int
579: \Phi_{\Gamma,X_z,X_{xy}} \Phi_h\,dz\,d\bbox{\rho}, \nonumber \\
580: &&{\cal P}_2=\sum_{\Gamma,X}U_{\Gamma,X}
581: \int [\Phi_{\Gamma,X_z,X_{xy}}
582: \varphi_h+ \Phi_h \varphi_{\Gamma,X_z,X_{xy}}]
583: \,dz\,d\bbox{\rho}, \\
584: &&{\cal P}_3=\sum_{\Gamma,X}U_{\Gamma,X}
585: \int \varphi_{\Gamma,X_z,X_{xy}} \varphi_h
586: \,dz\,d\bbox{\rho}. \nonumber
587: \end{eqnarray}
588: Here $\Phi=\Phi^l$, $\varphi=\varphi^l$ if $z<\xi$;
589: $\Phi=\Phi^r$, $\varphi=\varphi^r$ if $z>\xi$;
590: $U_\Gamma=\Omega_0^{-1}\int_{\Omega_0} u_\Gamma({\bf r})
591: \bbox{\nabla} v({\bf r})\,d{\bf r}$,
592: $U_X= \Omega_0^{-1}\int_{\Omega_0}u_{X}({\bf r})
593: \bbox{\nabla} v({\bf r})\,d{\bf r}$,
594: and $\Omega_0$ is the unit cell.
595:
596: The rate of the exciton recombination is
597: \begin{eqnarray}
598: \label{w}
599: &&w=\Lambda\left(
600: |{\cal P}_1|^2+{\cal P}_1{\cal P}_2^*+{\cal P}_1^*{\cal P}_2
601: +{\cal P}_1{\cal P}_3^*+{\cal P}_1^*{\cal P}_3
602: +|{\cal P}_2|^2\right), \nonumber\\
603: &&\Lambda=\frac{4\hbar e^2\omega}{3m_e^2c^3}G^2({\bf 0}).
604: \end{eqnarray}
605: Where $\hbar\omega$ is the exciton energy, $e$, $m_e$ and $c$ are
606: the fundamental constants.
607:
608: The luminescence magnitude $I(t)$ is
609: proportional to the recombination rate $w$ and the number of
610: excitons at the time $t$. We assume this number to be
611: proportional to $\exp(-wt)$ (or $\exp[-(w_0+w)t]$, if some
612: nonstochastic process with the rate $w_0$ occurs). The $w$
613: value is stochastical, since it depends on $\xi$. Therefore, to
614: determine the luminescence magnitude, we have to average the
615: value of $w\exp(-wt)$ over the realization of the random
616: function $\xi$. This could be done if we know the distribution
617: $P(w)$ of the $w$ value: $\overline{w\exp(-wt)}=\int_0^\infty
618: w\exp(-wt)P(w)\,dw$. The distribution $P(w)$ essentially
619: depends on ${\cal P}_1$, whether or not it vanishes.
620:
621: If ${\cal P}_1=0$ (i.e., if
622: the exciton recombination at a plane interface is forbidden) then
623: $w$ is proportional to squared module of
624: ${\cal P}_2$. The linear dependence between ${\cal P}_2$ and the
625: random variable $\xi$ follows from Eqs.~(\ref{11}) and
626: (\ref{14}). Therefore, if the distribution of $\xi$ is
627: Gaussian, then the distribution of ${\cal P}_2$ is also
628: Gaussian and the distribution of $w$ is exponential. This means
629: applicability of arguments of Refs.~\cite{Klein,Minami}, so that
630: $I(t)$ is determined by Eq.~(\ref{1}) where $w_r=\Lambda|{\cal
631: P}_2|^2$. For the case of $X_{xy}$ exciton we have
632: \begin{eqnarray*}
633: {\cal P}_2&=&\frac{4pqU_X}{\sqrt{N_1N_2}}
634: \sum_{\bf g}\left[
635: \frac{1-t_{22}^h}{\gamma_{X_{xy}}^2}
636: \tilde{\xi}\left({\bf q}_X+{\bf g}\right)
637: \right.\\
638: &+&\left.\frac{1-t_{22}^{xy}}{\gamma_h^2}
639: \tilde{\xi}^*\left({\bf q}_X+{\bf g}\right)
640: \right],
641: \end{eqnarray*}
642: and
643: \begin{equation}
644: \label{15}
645: w_r=\frac{16\pi^4a^4|U_X|^2\Lambda}{d_1^3d_2^3}
646: \left[
647: \frac{(1-t_{22}^h)^2}{\gamma_{X_{xy}}^4}+\frac{(1-t_{22}^{xy})^2}{\gamma_h^4}
648: \right]
649: \tilde{W}\left(\frac{\pi}{a}\right).
650: \end{equation}
651: Where $\tilde{W}({\bf k})=
652: \int W(\bbox{\rho})e^{-ik\rho}\,d^2\bbox{\rho}$
653: is the Fourier transform of
654: the correlation function; ${\bf q}_X=\left\{2\pi/a,0,0\right\}$
655: is the wave number of the $X$ valley, ${\bf g}$ is a
656: two-dimensional reciprocal lattice vector, it arises here since
657: integration in Eq.~(\ref{11}) has not been restricted by the
658: first Brillouin zone.
659:
660: If ${\cal P}_1\neq 0$ (i.e., if
661: the exciton recombination at a plane interface is allowed) then
662: the linear with respect to ${\cal P}_2$ terms in Eq.~(\ref{w})
663: are nonzero. This allows to omit the terms with ${\cal P}_3$
664: and $|{\cal P}_2|^2$, which are quadratic in $\xi$, or replace
665: them with their average values. Then $w$ becomes the linear
666: function of the random variable $\xi$. If the distribution of
667: $\xi$ is Gaussian then the distribution of $w$ is Gaussian too,
668: i.e.,
669: \[
670: P(w)=\frac{1}{\sigma\sqrt{2\pi}}e^{-\frac{(w-\overline{w})^2}{2\sigma^2}}.
671: \]
672: Where $\overline{w}=\Lambda|{\cal P}_1|^2$, and
673: $\sigma=\left[\overline{|w-\overline{w}|^2}\right]^{1/2}$.
674:
675:
676:
677: Hence
678: \begin{eqnarray}
679: \label{16}
680: I(t)&=&\frac{e^{-w_0t}}{\sigma\sqrt{2\pi}}
681: \int_0^\infty we^{-wt-\frac{(w-\overline{w})^2}{2\sigma^2}}\,dw
682: \\
683: &=&e^{-(\overline{w}+w_0)t}
684: \left[
685: \frac{\sigma}{\sqrt{2\pi}}+\frac{\overline{w}-\sigma^2t}{2}
686: e^{\frac{\sigma^2t^2}{2}}\mbox{erfc}\left(\frac{\sigma
687: t}{\sqrt{2}}
688: \right)
689: \right].
690: \nonumber
691: \end{eqnarray}
692:
693: If $|{\cal P}_1|\gg |{\cal P}_2|$, then
694: $\sigma^2\simeq
695: 2\Lambda^2
696: |{\cal P}_1|^2 \overline{{\cal P}_2{\cal P}_2^*}
697: $. For the case of $X_z$ exciton we have
698: \begin{eqnarray}
699: \label{16.5}
700: &&{\cal P}_1=
701: \frac{2a^3}{\sqrt{d_1d_2}}
702: \left[
703: 4\left(\frac{d_1}{d_2}\right)
704: \frac{t_{13}^z}{\gamma_{X_z}t_{44}^z}U_\Gamma+
705: \left(\frac{a}{d_1}\right)
706: \frac{1}{\gamma_{h}t_{22}^h}U_X
707: \right],\\
708: &&{\cal P}_2=
709: \frac{2a^2}{\sqrt{d_1d_2}}
710: \left[
711: 8\left(\frac{d_1}{d_2}\right)
712: \frac{{t_{13}^z}}{{t_{44}^z}}U_\Gamma+\right.\nonumber\\
713: &&\phantom{{\cal P}_2=}\left.
714: \frac{2\pi i a}{t_{22}^hd_1}
715: \left(
716: \frac{1}{t_{44}^z\gamma_{X_z}d_2}
717: -\frac{1-t_{22}^h}{t_{22}^h\gamma_hd_1}
718: \right)U_X
719: \right]
720: \eta(\xi)\tilde{\xi}(0),\nonumber
721: \end{eqnarray}
722: so that
723: \begin{eqnarray}
724: \label{17}
725: &&\overline{w}=
726: \frac{4\Lambda a^6}{d_1d_2}
727: \left[
728: 4\left(\frac{d_1}{d_2}\right)
729: \frac{t_{13}^z}{\gamma_{X_z}t_{44}^z}U_\Gamma+
730: \left(\frac{a}{d_1}\right)
731: \frac{1}{\gamma_{h}t_{22}^h}U_X
732: \right]^2,\\
733: &&\sigma^2=
734: \frac{2\Lambda a^4\overline{w}}{d_1d_2}
735: \left[
736: 64\left(\frac{d_1}{d_2}\right)^2
737: \frac{{t_{13}^z}^2}{{t_{44}^z}^2}U_\Gamma^2+
738: \right.
739: \nonumber\\
740: &&\left.\phantom{\sigma^2=}
741: \left(\frac{2\pi a}{t_{22}^hd_1}\right)^2
742: \left(
743: \frac{1}{t_{44}^z\gamma_{X_z}d_2}
744: -\frac{1-t_{22}^h}{t_{22}^h\gamma_hd_1}
745: \right)^2U_X^2
746: \right]
747: \tilde{W}(0).\nonumber
748: \end{eqnarray}
749: The first terms in square brackets can be interpreted as
750: a electron conversion from the $X$ valley of AlAs to
751: the $\Gamma$ valley of GaAs followed by the elecron-hole
752: recombination; they are small, since $t_{13}^z\ll 1$. The
753: second ones are due to indirect electron-hole
754: recombination;\cite{PR98} they occur only at the interface and,
755: therefore, have a small factor $a/d_{1,2}$. This factor is not
756: so small in short-period superlattices where $d_{1,2}$ are as
757: large as a few lattice constants. The indirect electron-hole
758: recombination prevails in such structures, if $a/d_{1,2}\gg
759: t_{13}^z$. We omit the terms that contain both these factors or
760: $|1-t_{44}^z|\ll 1$.
761:
762: The question arises, how small should be $|{\cal P}_1|$ in
763: order to Eq.~(\ref{1}) holds? This is possible if the
764: deviation of $|{\cal P}_2|^2$ from its average value in
765: Eq.~(\ref{w}) essentially exceeds $|{\cal P}_1{\cal P}_2^*|$,
766: i.e., when $|{\cal P}_1|^2\ll (h/l)^2|{\cal P}_2|^2$ or
767: \begin{equation}
768: \label{criterium}
769: \frac{\overline{w}^2}{\sigma^2}\ll
770: \frac{h^2}{l^2}.
771: \end{equation}
772:
773:
774: \section{Kinetics of exciton luminescence in type-II
775: G\symbol{"61}A\symbol{"73}/A\symbol{"6C}A\symbol{"73}
776: superlattices. Experiment}
777:
778: The undoped GaAs/AlAs type-II superlattices used in this study
779: were grown by molecular-beam-epitaxy at $600^o$C on a
780: (100) GaAs substrate. The sample BP205, where the
781: $X_z$ excitons were studied, contains 40 periods of 19.8-$\AA$
782: GaAs/25.5-$\AA$ AlAs. The $X_{xy}$ excitons were studied in the
783: sample BP354. It contains 25 periods of 25-$\AA$
784: GaAs/83.5-$\AA$ AlAs.
785:
786: The time-resolved photoluminescence of $X_z$ excitons was
787: excited by a YAG:Nd pulse laser with wavelength $532\, nm$, the
788: pulse duration was $0.15\,\mu s$. The N$_2$ laser with
789: wavelength $337\, nm$ and pulse duration $7\,ns$ was used to
790: investigate the time-resolved photoluminescence of $X_{xy}$
791: excitons.
792:
793: The luminescence was analyzed by a double grating
794: monochromator equipped with a photomultiplier. Lifetime
795: measurements were made by the time correlated single-photon
796: counting technique. The samples were immersed in liquid helium.
797:
798: Figures 2 and 3 present the experimental results on the exciton
799: decay rates in our samples.
800: Theoretical curves was derived from Eqs.~(\ref{1}), (\ref{16}).
801: The values of parameters $w_0=320\,c^{-1}$, $w_r=0.002\times
802: 10^6\,c^{-1}$, $\overline{w}=0.1\times 10^6\,c^{-1}$, and
803: $\sigma=0.61\times 10^6\,c^{-1}$ ensure the best fit with the
804: experiment. We see that Eq.~(\ref{1}) fits the experimental
805: data for the decay rates of $X_{xy}$ excitons in the sample
806: BP354, whereas Eq.~(\ref{16}) is more appropriate for $X_{z}$
807: excitons in the sample BP205. Note that the value of $w_0$,
808: which is associated with the phonon-assisted recombination, is
809: small in both curves. That is really the case at a low
810: temperature. Recombination of $X_{xy}$ excitons is considerably
811: slower than that of $X_z$ excitons. This means that the
812: interfaces in our samples are perfect enough to apply our theory
813: for interpretation of the experimental data.
814:
815: Expressions (\ref{15}) -- (\ref{17}) allow to estimate the
816: function $\tilde{W}(k)$ at the points $k=0$ and $k=2\pi/a$.
817: This is not sufficient to determine the function. However, it is
818: possible to estimate the parameters of the rough interface if we
819: restrict ourself to the particular type of the correlation
820: function. We assume the exponential correlation function
821: \begin{equation}
822: \label{18}
823: W(\rho)=h^2\exp(-\rho/l),
824: \end{equation}
825: where $l$
826: is the correlation length. This type of the correlation
827: function is more appropriate to our model of the rough interface
828: (Fig.~1); it allows to construct the two-position distribution,
829: so that the distribution of slopes has a $\delta$-singularity,
830: i.e., the slope is always zero exept a set of points (like point
831: 1) with measure zero.\cite{Berry}
832: This is impossible for the Gaussian correlation function
833: $W(\rho)=h^2\exp(-\rho^2/l^2)$ most frequently employed in
834: theoretical discussions.\cite{Kosobukin} Fourier
835: transform of the exponential function is
836: \begin{equation}
837: \label{18a}
838: \tilde{W}(k)= \frac{2\pi h^2l^2}{(1+k^2l^2)^{3/2}}.
839: \end{equation}
840: Unlike the Gaussian function it has not exponential factor,
841: which is small at a large $k$. This is also due to the
842: singular points 1; only in the vicinity of these points the
843: momentum relaxation of indirect $X_{xy}$ excitons is possible.
844:
845: If we assume that correlation functions are equal for the
846: interfaces of both our samples, then substitution of
847: Eq.~(\ref{18a}) into Eqs.~(\ref{15}) and (\ref{17}) allows to
848: find the values of $h$ and $l$.
849: The decay parameters of the wave functions
850: $\gamma_{X_z,X_{xy},h}$ in these
851: expressions are determined by Eq.~(\ref{11}) for the known
852: energies of the electrons or holes. As regards to
853: $U_{\Gamma,X}$, these values can be estimated only from the
854: band structure calculations for GaAs and AlAs. However, the
855: first terms in the expression for $\overline{w}$ and $\sigma^2$
856: (\ref{17}) can be omitted. Indeed, $t^z_{13}=t_{\Gamma
857: X}m^{\rm GaAs}_\Gamma/(m_ea\gamma^r)<0.06$, whereas
858: $a/d_{1}=2/7$, i.e., the indirect recombination of
859: $X_z$ exitons at the interface prevails in our samples. Then
860: the values of ${\sigma^2}/{\overline{w}^2}$ and
861: $w_r/{\overline{w}}$, which are determined from experimental
862: data, become independent of $U_{\Gamma,X}$. For our experiments
863: this estimation yields
864: \widetext
865: \begin{eqnarray}
866: \label{19}
867: &&\frac{\sigma^2}{\overline{w}^2}=
868: \frac{1}{2}\left(\frac{2\pi}{a}\right)^2
869: \left(
870: \frac{\gamma_h}{\gamma_{X_z}d_2}-
871: \frac{1-t^h_{22}}{t^h_{22}d_1}
872: \right)^2\tilde{W}(0),
873: \nonumber \\
874: &&\frac{w_r}{\overline{w}}=
875: \frac{4\pi^4\gamma_h^2{t^h_{22}}^2
876: {d_1^{X_{xy}}}^3d_2^{X_{xy}}}
877: {a^4{d_1^{X_z}}^3{d_2^{X_z}}^3}
878: \left[
879: \frac{(1-t_{22}^h)^2}{\tilde{\gamma}_{X_{xy}}^4}+
880: \frac{(1-t_{22}^{xy})^2}{\tilde{\gamma}_h^4}
881: \right]
882: \tilde{W}\left(\frac{2\pi}{a}\right).
883: \end{eqnarray}
884: Here $d_1^{X_z}$, $d_2^{X_z}$, $d_1^{X_{xy}}$, and
885: $d_2^{X_{xy}}$ are widths of GaAs and AlAs layers in the samples
886: BP205 ($d_1^{X_z}$, $d_2^{X_z}$) and BP354 ($d_1^{X_{xy}}$,
887: $d_2^{X_{xy}}$) where $X_z$ and $X_{xy}$ excitons were studied;
888: \begin{eqnarray*}
889: &&\gamma_{X_z}=\frac{1}{\hbar}
890: \sqrt{2m_{Xl}^{GaAs}
891: \left[E_X^{GaAs}-E_X^{AlAs}-\frac{\hbar^2}
892: {2m_{Xl}^{AlAs}(d_2^{X_z})^2} \right]},\\
893: &&\gamma_h=\frac{1}{\hbar}
894: \sqrt{2m_{hh}^{AlAs}
895: \left[E_h^{GaAs}-E_h^{AlAs}-\frac{\hbar^2}
896: {2m_{hh}^{GaAs}(d_1^{X_z})^2} \right]},
897: \quad
898: t_{22}^h=\frac{m_{hh}^{AlAs}}{m_{hh}^{GaAs}},\\
899: &&\tilde{\gamma}_{X_{xy}}=\frac{1}{\hbar}
900: \sqrt{2m_{Xt}^{GaAs}
901: \left[E_X^{GaAs}-E_X^{AlAs}-\frac{\hbar^2}
902: {2m_{Xt}^{AlAs}(d_2^{X_{xy}})^2} \right]},
903: \quad
904: t_{22}^{xy}=\frac{m_{Xt}^{AlAs}}{m_{Xt}^{GaAs}},
905: \\
906: &&\tilde{\gamma}_h=\frac{1}{\hbar}
907: \sqrt{2m_{hh}^{AlAs}
908: \left[E_h^{GaAs}-E_h^{AlAs}-\frac{\hbar^2}
909: {2m_{hh}^{GaAs}(d_1^{X_{xy}})^2} \right]},
910: \quad
911: \gamma^r=\frac{1}{\hbar}
912: \sqrt{2m_{\Gamma}^{AlAs}
913: \left(E_\Gamma^{AlAs}-E_X^{AlAs}\right)};
914: \end{eqnarray*}
915: \narrowtext
916: $E_{\Gamma,X,h}^{GaAs,AlAs}$ are positions of the band extrema
917: in GaAs and AlAs, $m_{\Gamma}^{GaAs}$ is effective mass of
918: $\Gamma$ valley of GaAs, $m_{Xl,Xt}^{GaAs,AlAs}$ are longitudial
919: and transversal effective masses in $X$ valleys of GaAs and
920: AlAs, $m_{hh}^{GaAs,AlAs}$ are effective masses of heavy holes
921: in GaAs and AlAs, and $m_e$ is mass of a free electron. We
922: assume $t_{21}^h\ll\gamma_h\sim 2/a$; this is the result of
923: calculations. \cite{Ando}
924:
925: Eq.~(\ref{19}) estimates the values of
926: For
927: the height $h$ and diameter $L$ [$L=4l$ for the distribution
928: (\ref{18})]\cite{Berry} of the roughnesses we find $h\approx
929: 1.25a$ and $L\approx 9a$. This is in agreement with structural
930: reseach of the GaAs/AlAs interface where the steps with the
931: height $h=a/2$ and the mean length of 40--200$\,\AA$ were
932: observed (see Ref.~\cite{Bechstedt} for the review).
933:
934: Rough estimation of $h$ and $l$ values also can be done if we
935: assume that criterium (\ref{criterium}) holds. This justifies
936: Eq.~(\ref{1}) for $X_z$ excitons where $w_0\equiv
937: w_0^{X_z}=\Lambda|{\cal P}_1|^2$ and $w_r\equiv
938: w_r^{X_z}=\Lambda|{\cal P}_2|^2$. Using Eq.~(\ref{16.5}), we
939: find the expressions for $2w_r^{X_z}/w_0^{X_z}$ and
940: $w_r/w_0^{X_z}$ [unlike $w_r^{X_z}$ the $w_r$ value
941: correspondes to $X_{xy}$ excitons (Fig.~3) and determined by
942: Eq.~(\ref{15})]. These expressions accept the form of
943: Eq.~(\ref{19}) after the substitutions
944: $\sigma^2/\overline{w}^2\rightarrow 2w_r^{X_z}/w_0^{X_z}$ and
945: $w_r/\overline{w}\rightarrow w_r/w_0^{X_z}$ of their left sides.
946: The values of $w_0^{X_z}=0.11\times 10^6\,c^{-1}$ and
947: $w_r^{X_z}=0.38\times 10^6\,c^{-1}$ ensure the best fit of the
948: dashed line (Fig.~2) with experiment. This estimation yields
949: $h=a$, $L=8.8a$, which are close to the values obtained from
950: Eq.~(\ref{16}). For this reason both theoretical curves (Fig.~2)
951: fit experimental data at small times. Nevertheless,
952: Eq.~(\ref{16}) better fits the experimental data at large times
953: where it unsure the slower decay of the luminescence:
954: $I(t)\propto \exp (-\overline{w}t)/t$, instead of $I(t)\propto
955: \exp (-w_0t)/t^{3/2}$ as it is predicted by Eq.~(\ref{1}).
956:
957:
958:
959:
960:
961: \section{Discussion}
962:
963: In this paper we investigate the exciton luminescence in type II
964: GaAs/AlAs superlattices. We use the envelope function
965: approximation to consider the exciton recombination at an
966: interface. To justify this approach, we have to note that
967: envelope function approximation has been used only to find the
968: reflection and transmission coefficients. While the Bloch
969: functions $f_e$ and $f_h$ has been used to find the probability
970: of the exciton recombination. The error arises only when we
971: consider the Bloch amplidudes $u_\Gamma({\bf r})$, $u_X({\bf
972: r})$, and $v({\bf r})$ as periodical functions at the interface.
973: Indeed, the deviation of these amplitudes from their bulk values
974: arises only at a small distance from the interface; this
975: deviation is especially small for the contacts of similar
976: materials (e.g., GaAs/AlAs).\cite{Ando,Allmen}
977:
978: It seems the boundary conditions (\ref{bcplane}) connect a
979: very few valleys of the electron spectrum to consider the
980: interface influence on the exciton recombination; nevertheless,
981: it is not the case. Indeed, the electron wave functions in the
982: valleys that are not explicitly involved in Eq.~(\ref{bcplane}) are
983: strongly localized at the interface. This allows to consider
984: them in terms of the boundary conditions where the parameters
985: $t_{ik}$ are influenced by these valleys. This procedure had
986: been described when Eq.~(\ref{bcplane}) was derived. The error arises
987: only when these parameters are considered as independent of the
988: electron energy; that is possible if the energy difference
989: between the bottoms of the appropriate valleys considerably
990: exceeds the exciton energy. Note that a lot (about 10) of the
991: electron bands are sometimes taken into account when the
992: parameters of the interface matrix
993: are calculated.\cite{Grinyaev}
994:
995:
996: We use the boundary conditions for the envelope wave function
997: to consider $\Gamma-X$ mixing of electrons at
998: the interface. This approach is more general than the kinetic
999: model proposed in Ref.~\cite{Maaref}. The kinetic equation where
1000: the electron states in the $\Gamma$ and $X$ valleys are
1001: considered as independent can be used for low $\Gamma-X$
1002: mixing. Only in that case it is possible to add the
1003: probabilities for the electron to be in $\Gamma$ and $X$
1004: valleys. It should be noted that we also assume the small value
1005: of $\Gamma-X$ mixing ($|t_{13}^z|\ll 1)$. However, this
1006: approximation is not principal for our consideration; it only
1007: makes the results [Eqs.~(\ref{11}), (\ref{15}), (\ref{17}), and
1008: (\ref{19})] not so cumbersome.
1009:
1010: Influence of a nonstochastic process on the exciton recombination
1011: in Ref.~\cite{Klein} is considered by the exponential factor
1012: $e^{-w_0t}$. This factor could be obtained if we
1013: insert the corresponding term in $\tau$-approximation
1014: into the kinetic equation for the exciton density. If the
1015: $\tau$-approximation is not applicable for the process, then
1016: this factor becomes nonexponential.\cite{Krivirotov} Correlation
1017: between stochastic and nonstochastic processes changes the
1018: second factor in Eq.~(\ref{1}). In this case the probability of
1019: the exciton recombination $w$ is not a simple sum of the
1020: probabilities of each process. As the result, the additional
1021: terms arise in the expression for $w$ [the second and third
1022: terms in Eq.~(\ref{w})]. These terms are linear in the
1023: stochastic variable, so that their averages vanish. For this
1024: reason they are not important when the mean intensity of the
1025: luminescence or the light absorption\cite{PhysicaE} is
1026: considered. However, they are important for the kinetic
1027: phenomena, because they determine the mean square of the
1028: deviation $\sigma$ of the stochastic variable from its mean
1029: value. The nonexponential behavior of the decay rate
1030: Eq.~(\ref{16}) valids any time when linear with respect to
1031: the stochastic variable terms are main in
1032: the expression for $w$. Such a situation can occur also in
1033: other type-II semiconductor structures where the interface
1034: influence is essential, e.g., in quantum dots.\cite{Govorov}
1035:
1036: Expressions (\ref{15}) and (\ref{17}) relate parameters of the
1037: radiative decay rates ($w_r, \overline{w}$, and $\sigma$) with
1038: the correlation function of the rough interface. The values of
1039: the Fourier transform of this function at two particular points,
1040: $k=0$ and $k=2\pi/a$, are necessary for these relations. This
1041: allows to estimate the parameters only for simplest
1042: functions [like Eq.~\ref{18}]. The real interface might be more
1043: complicated. In particular, a few different
1044: scales could be characteristic for the roughnesses at the interface.
1045: Expressions (\ref{15}) and (\ref{17}) takes into account all
1046: these factors; however, it is impossible to determine more than
1047: two parameters from the time-resolved luminescence experiments
1048: only.
1049:
1050: Comparing the experimental results (Figs. 2 and 3), we see that
1051: mean lifetime of $X_{xy}$ excitons essentially exceeds that of
1052: $X_z$ excitons. This happens due to
1053: recombination of $X_z$ excitons at a plane interface.
1054: Meanwhile, influence of the roughnesses, i.e., the
1055: nonexponential factor in $I(t)$, is more essential for $X_z$
1056: excitons. This can be understood from our analysis. Indeed,
1057: $\sigma\propto \tilde{W}(0)$, whereas $w_r\propto
1058: \tilde{W}(2\pi/a)$ while $\tilde{W}(2\pi/a)\ll \tilde{W}(0)$.
1059: The recombination occurs in some region near the step (point 1
1060: in Fig.~1). The size of this region is of the order of
1061: $|{\bf q}_\parallel|^{-1}$, where ${\bf q}_\parallel$ is the
1062: parallel-to-interface component of the electron wave vector.
1063: This region is large for $X_z$ electrons ($|{\bf
1064: q}_\parallel|\simeq r_B^{-1}$, where $r_B$ is the exciton
1065: radius) but it is small for $X_{xy}$ electrons ($|{\bf
1066: q}_\parallel|\simeq 2\pi/a$). As the result, the small factor
1067: [of the order of $(a/l)^3$] arises in the expression for $w_r$.
1068:
1069: In conclusion, kinetics of the exciton luminescence
1070: at a rough interface has been considered. The Klein {\it at al.}
1071: law (\ref{1}) is shown to be valid for the decay rate of
1072: $X_{xy}$ excitons, whereas the more complicated expression
1073: (\ref{16}) is applicable for $X_z$ excitons. Expressions
1074: (\ref{15}) and (\ref{17}), which relate the parameters of the
1075: exciton kinetics with statistical characteristics of the rough
1076: interface, allow to estimate some of these characteristics from
1077: the experimenal data. The values of the mean height $7\,\AA$
1078: and length $50\,\AA$ of the roughnesses obtained from
1079: our experiments are in a good agreement with the results of
1080: structural investigations of the GaAs/AlAs interface.
1081:
1082: \acknowledgments
1083: Authors wish to thank
1084: Prof.\ E.\ L.\ Ivchenko for valuable discussions. This work
1085: was supported by the Russian Foundation for the Basic Research,
1086: Grants No.~99-02-17019, 98-02-17896, 00-02-17658, and the
1087: Program ''Physics of Solid State Nanostructures'' by the Russian
1088: Interdisciplinary Scintific and Technical Council, Grant
1089: No.~99-1133.
1090:
1091:
1092:
1093: \begin{references}
1094: \bibitem{Klein}M.~V.~Klein, M.~D.~Sturge, and E.~Cohen, Phys.
1095: Rev. B, {\bf 25}, 4331 (1982).
1096:
1097: \bibitem{Minami}F.~Minami {\it et al.}, Phys. Rev. B, {\bf 36},
1098: 2875 (1987).
1099:
1100: \bibitem{Krivirotov}I.~N.~Krivorotov {\it et al.}, Phys. Rev. B,
1101: {\bf 58}, 10687 (1998).
1102:
1103: \bibitem{Nagao}S.~Nagao {\it et al.}, J. of Crystal Growth,
1104: {\bf 175}, 10687 (1997);
1105: B.~A.~Wilson {\it et al.}, Phys. Rev. B, {\bf 40}, 1825
1106: (1989);
1107: E.~Finkman {\it et al.}, J.~Lumin. {\bf 39}, 57 (1987);
1108: B.~A.~Wilson {\it et al.}, J. Vac. Sci. Technol. B {\bf 6}, 1156
1109: (1988).
1110:
1111:
1112:
1113: \bibitem{Ivchenko}E.\ L.\ Ivchenko and G.\ E.\ Pikus, {\it
1114: Superlattices and Other Heterostructures. Symmetry and Optical
1115: Phenomena}, second ed. (Springer-Verlag, Berlin, 1997).
1116:
1117:
1118:
1119: \bibitem{PhysicaE}L.~Braginsky, Physica E, {\bf 5}, 142 (1999).
1120:
1121: \bibitem{Ando}T.~Ando and H.~Akera,
1122: Phys. Rev. B, {\bf 40}, 11619 (1989).
1123:
1124: \bibitem{Lurssen}D.~L\"uer\ss en {\it et al.}, Phys. Rev. B, {\bf 59},
1125: 15862 (1999).
1126:
1127: \bibitem{Bechstedt}F.~Bechstedt, R.~Enderlein, {\it
1128: Semiconductor Surfaces and Interfaces. Their Atomic and
1129: Electronic Structures}, (Berlin: Akademie-Verlag, 1988).
1130:
1131: \bibitem{Ivchenko1}I.~L.~Aleiner and E.~L.~Ivchenko, Fiz. Tech.
1132: Poluprovodn. {\bf 27} 594 (1993) [Sov. Phys. Semicond. {\bf 27}
1133: (1993)]; Y.~Fu, M.~Willander, E.~L.~Ivchenko, and A.~A.~Kiselev,
1134: Phys. Rev. B, {\bf 47} 13498 (1993).
1135:
1136:
1137: \bibitem{Bass}F.\ G.\ Bass and I.\ M.\ Fuks, {\it Wave
1138: Scattering from Statistically Rough Surfaces}, (Pergamon Press,
1139: 1979).
1140:
1141: \bibitem{PR98}L.~Braginsky, Phys. Rev. B, {\bf 57}, R6870 (1998).
1142:
1143:
1144: \bibitem{Berry}M.~V.~Berry, Phil. Trans. {\bf A273}, 611 (1973);
1145: J.~M.~Ziman, {\it Models of Disorder. The theoretical physics of
1146: homogeneously disordered systems}, (Cambrige University Press,
1147: 1979).
1148:
1149: \bibitem{Kosobukin}For the exciton problem this approximation
1150: has been used in V.~A.~Kosobukin, Fiz. Tverd. Tela
1151: {\bf 41}, 330 (1999) [ Phys. of Solid States {\bf 41}, XXX
1152: (1999)]; see also the references therein.
1153:
1154:
1155:
1156: \bibitem{Allmen} P.~von~Allmen, Phys. Rev. B, {\bf 46}, 15377
1157: (1992).
1158:
1159: \bibitem{Maaref}M.~Maaref {\it et al.}, Solid State Commun.,
1160: {\bf 81}, 35 (1992).
1161:
1162: \bibitem{Grinyaev}S.~N.~Grinyaev, private communication; see,
1163: e.g., S.~N.~Grinyaev and G.~F.~Karavaev, Fiz. Tverd. Tela
1164: {\bf 42}, 752 (2000) [ Phys. of Solid States {\bf 42}, XXX
1165: (2000)].
1166:
1167: \bibitem{Govorov} A.~O.~Govorov and A.~V.~Chaplik, Sov. Phys.
1168: JETP 72, 1037 (1991) [Zh. Eks. Teor. Fiz. 99, 1853 (1991)];
1169: A.~V.~Kalameitsev, A.~O.~Govorov, and V.~Kovalev, JETP Lett.
1170: 68, 669 (1998) [Pis'ma Zh. Eks. Teor. Fiz. 68, 634 (1998)].
1171:
1172: \end{references}
1173:
1174: \begin{figure}
1175: \caption{
1176: The model of the rough interface: side view.
1177: }
1178: %\label{}
1179: \end{figure}
1180:
1181: \begin{figure}
1182: \caption{
1183: Temporal evolution of the $X_z$ exciton emmision.
1184: Theoretical curves (dashed and solid lines) was derived from
1185: Eqs.~(\ref{1}) and (\ref{16}) respectively. Dotts show
1186: the experimental data. } %\label{}
1187: \end{figure}
1188:
1189: \begin{figure}
1190: \caption{
1191: Temporal evolution of the $X_{xy}$ exciton emmision.
1192: Theoretical curve (solid line) was derived from
1193: Eq.~(\ref{1}). Dotts show the
1194: experimental data.
1195: }
1196: %\label{}
1197: \end{figure}
1198:
1199:
1200: \end{document}
1201:
1202: