cond-mat0009390/pt.tex
1: %\documentstyle[prb,aps,psfig]{revtex}
2: %\documentstyle[preprint,aps,psfig]{revtex}
3: \documentstyle[prb,aps,psfig,twocolumn]{revtex}
4: \begin{document} 
5: \title{Role of bound pairs in the optical properties of highly excited
6: semiconductors: a self consistent ladder approximation approach.}
7: 
8: \twocolumn[\hsize\textwidth\columnwidth\hsize\csname@twocolumnfalse\endcsname
9: \author{C. Piermarocchi} \address{Department of Physics, University of
10: California San Diego, La Jolla, CA 92093-0319} \author{F. Tassone}
11: \address{Scuola Internazionale Superiore di Studi Avanzati, via Beirut
12: 4, I-34014 Trieste, Italy} \date{\today} \maketitle %\draft
13: 
14: 
15: 
16: \begin{abstract}
17: Presence of bound pairs (excitons) in a low-temperature electron-hole
18: plasma is accounted for by including correlation between fermions at
19: the ladder level. Using a simplified one-dimensional model with
20: on-site Coulomb interaction, we calculate the one-particle
21: self-energies, chemical potential, and optical response.  The results
22: are compared to those obtained in the Born approximation, which does
23: not account for bound pairs.  In the self-consistent ladder
24: approximation the self-energy and spectral function show a
25: characteristic correlation peak at the exciton energy for low
26: temperature and density. In this regime the Born approximation
27: overestimates the chemical potential.  Provided the appropriate vertex
28: correction in the interaction with the photon is included, both ladder
29: and Born approximations reproduce the excitonic and free pair optical
30: absorption at low density, and the disappearance of the exciton
31: absorption peak at larger density.  However, lineshapes and energy
32: shifts with density of the absorption and photoluminescence peaks are
33: drastically different.  In particular, the photoluminescence emission
34: peak is much more stable in the ladder approximation. At low
35: temperature and density a sizeable optical gain is produced in both
36: approximations just below the excitonic peak, however this gain shows
37: unphysical features in the Born approximation. We conclude that at low
38: density and temperature it is fundamental to take into account the
39: existence of bound pairs in the electron-hole plasma for the
40: calculation of its optical and thermodynamic properties.  Other
41: approximations that fail to do so are intrinsically unphysical in this
42: regime, and for example are not suitable to address the problem of
43: excitonic lasing.
44: 
45: \end{abstract} \pacs{}
46: 
47: ]
48: 
49: 
50: %\begin{multicols}{2}
51: 
52: 
53: \section{Introduction} 
54: 
55: The binding of a gas of oppositely charged fermions into a bosonic gas
56: of bound pairs has a fundamental interest in condensed matter physics,
57: in problems ranging from the state of hydrogen to the nature of the
58: electron-hole plasma in Si and Ge. This pairing takes its origin from
59: the Coulomb correlation between the charged fermions, and strongly
60: affects the statistical and thermodynamic properties of these systems.
61: In semiconductor physics, the role of the Coulomb correlation and
62: pairing in the description of the electron-hole plasma in Si and Ge
63: has been extensively investigated. Refined descriptions of both the
64: ground state and the thermodynamic properties of this system have been
65: developed in the past thirty years. In particular, an excitonic
66: insulator ground state was first proposed and then shown to be less
67: stable than a simpler electron-hole liquid in Si and Ge, due to
68: band-structure effects.\cite{rice77} Thermodynamics of the
69: electron-hole gas in simpler crystals was also addressed by
70: Haug\cite{haug76}, and Zhu {\it et al.}\cite{zhu96} aiming at deriving
71: a complete phase diagram from variational approaches.
72: 
73: 
74: 
75: 
76: In this paper we will deal with direct gap semiconductors and we will
77: discuss the thermodynamics of the photoexcited electron and holes at
78: quasi-thermal equilibrium. Our main concern is to determine how the
79: inclusion of pairing effects in the theoretical description, as done
80: for the systems mentioned above, can affect predictions on the shape
81: of optical spectra. In the optical properties of semiconductors, the
82: effect of the Coulomb electron-hole correlation comes into play to
83: explain the characteristic excitonic absorption and emission. These
84: excitonic features can be described by including vertex corrections in
85: the interaction of the electron-hole pair with the
86: photon.\cite{elliott57} In particular, excitonic absorption is found
87: in an empty crystal, when a background electron-hole plasma is absent.
88: However, when carriers are electrically or optically injected, they
89: reach quasi-thermal equilibrium through scattering, and a background
90: electron-hole plasma is eventually formed. In this typical situation
91: the absorption is modified, and spontaneous emission
92: (photoluminescence) also takes place. Absorption and emission
93: give valuable information about the state of the plasma, but a theory
94: describing both the thermodynamics and optical properties of the
95: plasma is required in order to extract this information.
96: Semiconductor Bloch equations, and their evolutions, are certainly
97: among the most known and used theories for this
98: purpose.\cite{haugkoch} In these theories, the electron-hole plasma
99: has been originally treated at the Hartree-Fock level,\cite{haug84}
100: and eventually screening of the Coulomb interaction was also included
101: at various levels of approximation.\cite{haugkoch} However, none of
102: these refinements accounts for the existence of bound excitons in the
103: plasma. This is certainly a weak point at low temperatures, when
104: simple thermodynamic arguments show that condensation of electron-hole
105: gas into a gas of bound-electron hole pairs is favorable in a wide
106: range of densities.
107: 
108: 
109: In the following sections we will show the relevant changes in the optical
110: properties of the semiconductor when the existence of excitons at low
111: temperatures is taken into account. This is clearly a leap beyond the
112: standard description of the interaction of light with the
113: semiconductor as described above.  Our purpose is qualitative, but we
114: can still gain solid understanding of the underlying physics. In our
115: opinion at this stage of development it is not wise to aim at
116: improving also quantitatively existing theories which are already far
117: too involved.  We keep the problem simple enough so as to always control
118: and understand the resulting physics. We thus consider an accurately
119: solvable model, where the system is one-dimensional, Coulomb
120: interaction is on-site, and electron and holes are spin-polarized.
121: Excitonic correlation is also expected to be stronger in one dimension
122: than higher dimensions. Moreover, for the single-pair problem, exactly
123: a single bound state is obtained with the on-site interaction.  In
124: our description of the electron-hole plasma, we either include carrier
125: scattering at the Born level, which does not describe bound pairs in
126: the plasma, or we include enough correlation so as to describe the
127: pairing into bound states. For this purpose, ladder diagrams in the
128: electron-hole scattering kernel\cite{kraeft86,zimmermann88,pereira98}
129: are included, and the electron and hole self energies calculated with
130: it.\cite{tassone99} The resulting Dyson equations for the Green
131: functions are solved self-consistently. The approximation is therefore
132: called self-consistent ladder approximation (SCLA). Screening is
133: expected to be weak in one dimension,\cite{benner91} due to reduced
134: screening phase-space, and we neglect it altogether.  We remark that
135: in the SCLA we calculate the single particle propagators and
136: self-energies only, thus in this sense, the theory remains purely
137: fermionic. We do not map the model onto bound and un-bound states at
138: any time. However, it is possible to show analytically that at low
139: temperatures and densities, a bosonic gas of excitons is effectively
140: described, and that self-consistency effectively introduces scattering
141: between these excitons. When this analogy is extended to larger
142: densities and temperatures, where bound and free carriers are expected
143: to coexist, we understand that self-consistency also accounts for
144: multiple exciton-free carrier, and free carrier-free carrier
145: scattering. Thus, a large amount of correlation, well beyond that
146: described in simpler Born approximations is included in the SCLA.  We
147: remark that we also assume quasi-thermal equilibrium in the electron-hole
148: gas, which is well justified as the characteristic radiative
149: recombination (several hundreds of picoseconds) is much slower than
150: the typical scattering time in the density range considered. The
151: theory is however written so as to be readily extended to the
152: non-equilibrium situation. We finally show how to correctly calculate
153: photon absorption and emission (photoluminescence or PL) in a
154: conserving sense, i.e. respecting f-sum rules. We only neglect
155: polaritonic effects, which are very weak in one
156: dimension.\cite{tassone.93}
157: 
158: 
159: 
160: 
161: Within the considered models, we highlight important trends in optical
162: and thermodynamical properties.  First, Born approximations
163: overestimate the chemical potential at low temperatures when a
164: consistent fraction of pairs is bound into excitons. Second, energetic
165: stability of the excitonic emission peak as a function of density is
166: better described in the SCLA, even well beyond disappearance of
167: excitonic absorption peak at large excitation densities. Third,
168: the large excitonic {\em gain} at densities just below those
169: where excitonic absorption disappears is clearly incorrectly described
170: in simpler Born approximations.  As these trends have a clear physical
171: origin, they are expected to hold qualitatively even when more
172: realistic interaction potentials (eventually including screening) are
173: considered. We conclude that these results do have important
174: experimental and theoretical relevance, signaling the necessity of an
175: adequate description of the excited semiconductor in the
176: low-temperature region. For example, current description of the
177: electron-hole plasma in the semiconductor Bloch equations is clearly
178: insufficient to address excitonic (inversion-less) lasing.
179: 
180: 
181: 
182: The paper is organized as follows. In Sec. \ref{sec:method} we
183: introduce and explain how to implement the self consistent ladder
184: approximation, and the Markov-Born approximation and self-consistent
185: Born approximation.  Single particle properties are analyzed in
186: Sec. \ref{sec:single}, where we show how excitonic correlation appears
187: in the electron and hole spectral functions.  Chemical potentials and
188: self-energies are then compared. In Sec. \ref{sec:excpot} we
189: analytically show how the SCLA at low temperature and density
190: describes an interacting exciton gas. Optical properties of a
191: semiconductor quantum wire are calculated in Sec. \ref{sec:optical}
192: for the three approximations. Absorption and emission as functions of
193: the density are also presented, and discussed in this
194: section. Conclusions are drawn is Sec. \ref{sec:conclusions}.
195: 
196: 
197: 
198: \section{Born and self consistent ladder approximation}
199: \label{sec:method}
200: 
201: 
202: In this section we introduce the SCLA, the simpler Markov-Born
203: approximation (MBA) and self-consistent Born approximation (SCBA). We
204: assume thermal equilibrium in the electron-hole plasma, but, in order
205: to avoid analytic continuation problems, we do not resort to the
206: Matsubara Green function technique,\cite{mahan81} which becomes
207: numerically delicate when the spectral function shows a large number
208: of poles. Instead, we work directly in the real time/frequency
209: space. This has the additional advantage that the approach can be
210: readily extended to non-equilibrium.\cite{danielewicz} We thus
211: consider four single particle Green's Functions:\cite{haug96} the
212: retarded, advanced, lesser and greater Green's functions defined as
213: \begin{mathletters}
214: \label{eq:greens}
215: \begin{eqnarray} 
216: G^+(1,2)&=&-i \theta(t_1-t_2) \langle T [c(1) c^\dagger(2)]_+ \rangle,
217: \label{eq:g+} \\ G^-(1,2)&=&i \theta(t_2-t_1) \langle T [c(1)
218: c^\dagger(2)]_+ \rangle, \label{eq:g-}\\ G^<(1,2)&=&i \langle
219: c^{\dagger}(2) c(1) \rangle, \label{eq:g<}\\ G^>(1,2)&=&-i \langle
220: c(1) c^{\dagger}(2)\rangle . \label{eq:g>}
221: \end{eqnarray}
222: \end{mathletters}
223: Here $T$ is the time ordering operator, $1=(x_1,t_1,\sigma_1)$, where
224: $x_1$ is the position, $t_1$ the time, and $\sigma_1$ is the spin
225: index. $c(1)$ is the electron annihilation operator and $[~ ]_+$ is
226: the anticommutator.  Similar expressions for the hole Green's
227: functions hold, with $ d(1)$ and $ d^{\dagger}(1)$ the hole
228: annihilation and creation respectively.  Only two of the four
229: functions above are independent, and in particular, we will consider
230: the retarded and lesser (or correlation) functions. Formally, a single
231: Green's function $G(x_1,\tau_1,x_2,\tau_2)$, may be introduced for
232: compactness of notation, with $\tau_1$ and $\tau_2$ defined on the
233: Keldysh contour, having the two branches shown in
234: Fig. \ref{fig:keldysh}. Following simple rules, any equation with
235: times defined on the Keldysh contour can be written back in a set of
236: equations defined on the ordinary time axis for the above four Green
237: functions (see for instance Ref. \onlinecite{haug96}).
238: \begin{figure}
239: \centerline{\psfig{figure=ptfig1.eps,width=8truecm}}
240: \vspace{1cm}
241: \caption{The Keldysh contour. Upper branch, normal time ordering,
242: lower branch, reversed time ordering.}
243: \label{fig:keldysh}
244: \end{figure}
245: The statistical averages are done in the macro-canonical ensemble:
246: $\langle \dots \rangle={\rm Tr}\{\rho \dots\}$ with $\rho=
247: e^{-\beta(H-\mu_{e} N_e -\mu_{h} N_h)}/{\rm Tr} \{e^{-\beta(H-\mu_{e}
248: N_e-\mu_{h} N_h)}\}$.  Here $\beta=1/T$ and $H$ is the total
249: Hamiltonian of the interacting electron-hole system.  In principle,
250: electrons and holes have independent chemical potentials, and
251: densities. However, we are interested in the description of a laser
252: excited semiconductor where the density of electrons and holes is the
253: same, and charge is balanced.  We will also assume that the
254: temperature of the two components of the interacting gas is the
255: same. In fact, after excitation, thermalization of carriers is mainly
256: driven by exchange of energy with the phonon thermal bath which, at
257: equilibrium, leads to $T_e = T_h = T $.  Moreover, for simplicity we
258: consider the same electron and hole masses, giving $\mu_e=\mu_h=\mu$
259: and the same Green functions and self-energies. Extension of the
260: theory to different masses is straightforward. Finally, the Green
261: functions above depend only on the time difference $t_2-t_1$ as the
262: system is stationary, and on the relative distance $x_2-x_1$ as the
263: system is homogeneous.
264: 
265:  In this thermal regime, the retarded and lesser Green functions
266: defined above are also related by the Kubo Martin Schwinger relations
267: \cite{kadanoff62}
268: \begin{mathletters}
269: \label{eq:KMSfermi}
270: \begin{eqnarray}
271: G^<(k,\omega)=-2 \Im [G^+(k,\omega)] f(\omega-\mu)\\ G^>(k,\omega)=2
272: \Im [G^+(k,\omega)](1-f(\omega-\mu)),
273: \end{eqnarray}
274: \end{mathletters}
275: where we considered the Fourier transforms with respect to the
276: relative time $t_1-t_2$, and with respect to the relative position
277: $x_1-x_2$. The function $f(\omega)=[exp(\omega/T)+1]^{-1}$ is the
278: Fermi function. In the stationary case
279: $G^{+}(k,\omega,\sigma)={G^{-}}^*(k,\omega,\sigma)$ contain the
280: information on the spectral properties of the quasiparticles, given by
281: the one particle spectral function $A(k,\omega,\sigma)=-2 \Im
282: [G^+(k,\omega,\sigma)]$.  The correlation functions instead contain
283: information on the quasi particle occupation number
284: $G^<(k,t_1=t_2=t,\sigma)=N(k,t,\sigma)$, and
285: $G^>(k,t_1=t_2=t,\sigma)=1-N(k,t,\sigma)$, where $N(k,t\sigma)$ are
286: the occupation numbers.
287: 
288: The on-site Coulomb interaction reads:
289: \begin{equation}
290: V(x) = \pm a \delta(x)~,
291: \label{eq:delta}
292: \end{equation} 
293: the potential is repulsive for electron-electron and hole-hole, and
294: attractive for electron-hole interaction. An attractive delta-like
295: potential between the electrons and the holes gives one bound state
296: only for the pair problem, and $a$ can be chosen such to reproduce the
297: exciton binding energy measured in the limit of low density.  The
298: constant $a$ has the dimensions of $energy \times length$, the exciton
299: Bohr radius is given by $a_B=\hbar^2/(m a)$ and the exciton binding
300: energy is $E_b=m a^2/2 \hbar^2$ where $m$ is the reduced mass of the
301: electron-hole system. We will use the units $\hbar=a=m=1$ throughout
302: the paper. In these units $a_B=1$ and $E_b=0.5$. We remind that
303: typical values for semiconductor quantum wires are $a_B=10^{-6}$ cm
304: and $E_b=10$ meV.\cite{optwires} With the above effective potential of
305: Eq. (\ref{eq:delta}) the interaction Hamiltonian is
306: $$H_C = \frac{1}{2}\sum_{1} c^\dagger (1)c^\dagger(1')c(1')c(1)+$$
307: $$+ \frac{1}{2}\sum_{1} d^\dagger (1)d^\dagger(1')d(1')d(1)+$$
308: \begin{equation}
309: -\sum_{1}
310: c^\dagger(1)d^\dagger(1')
311: c(1)d(1')~.\label{eq:H}
312: \end{equation} 
313: Here $1'=(x_1,t_1,\sigma'\ne \sigma)$ for the electron-electron and
314: hole hole interaction, while $1'=(x_1,t_1,\sigma')$ for the
315: electron-hole interaction, as due to the fermionic nature of the
316: carriers, only interaction with opposite spin is allowed in the
317: intraband term for a contact potential.  In the following, we assume a
318: spin-polarized system of electrons and holes, so that we keep only the last
319: term describing the electron-hole interaction in the Hamiltonian
320: (\ref{eq:H}). Generalization to both spins is straightforward.
321: 
322: 
323:  
324: The Dyson equation for the single particle $G(1,2)$ propagators read
325: \begin{equation}
326: G(1,2)=G_0(1,2)+\int
327: d\bar 3~d\bar 4 G_0(1,\bar 3)\Sigma(\bar 3,\bar 4)G(\bar 4,2)~, $$
328: \label{eq:dyson}
329: \end{equation}
330: where time integrations are over the Keldysh contour.  $G_0$ is the
331: free propagator. The functional dependence of the self energy $\Sigma$
332: on the single particle Green's functions determines the degree of
333: approximation. In particular, it can be expressed through an
334: electron-hole scattering kernel $T(1,2;1',2')$:
335: \begin{equation}
336: \Sigma(1,2)= i \int d\bar 1'~d\bar 2'~T(1,\bar 1';2,\bar 2') G(\bar
337: 2',\bar 1') \label{eq:sigma}
338: \end{equation}
339: In the Born approximation,
340: $$T(1,1';2,2')=V(1,1')+$$
341: \begin{equation} +
342:  V(1,1')i G(1,2)G(1',2') V(2,2'), $$
343: \label{eq:tborn}
344: \end{equation}
345: where $V(1,1')= V(x_1-x_{1'}) \delta(t_1,t_{1'})$, and $\delta(\tau,\tau')$
346: is the Dirac delta function extended on the Keldysh contour.  The
347: self-energy $\Sigma(k,\omega)$ can be calculated at the single
348: particle pole $\omega = k^2/4 $. In this case we are neglecting memory
349: terms in the scattering and thus using a Markov approximation; this
350: corresponds to the MBA. In the SCBA, the same scattering kernel of
351: Eq. \ref{eq:tborn} is used, but the full energy dependent structure of
352: the resulting self-energy given in Eq. (\ref{eq:sigma}) is used in the
353: Dyson equations Eq. (\ref{eq:dyson}).  The problem is then solved
354: self-consistently, as explained later. In this case, memory effects,
355: beyond the Markov approximation, are also included.  In the SCLA, the
356: kernel $T(1,1';2,2')$ is the solution of the Bethe Salpeter equations
357: in the ladder approximation, which for a generic potential reads
358: $$T(1,1';2,2')= V(1,1') +$$
359: \begin{equation} 
360: +\int d\bar 3~d\bar 3'~V(1,1')i G(1,\bar 3)G(1',\bar 3') T(\bar 3,\bar 3';2,2')
361: \label{eq:bethesalpetereh}
362: \end{equation} 
363: For a $\delta$-like potential, $V(1,1')=-\delta(1,1')$ and $T$ depends
364: only on $2-1$. The reduced equation reads:
365: \begin{equation} 
366: T(1;2)= -1 - i \int d{\bar 3}~G(1,\bar 3)G(1,\bar 3) T (\bar 3;2)
367: \label{eq:bethesalpetereh2}
368: \end{equation}
369:  The Eqs. (\ref{eq:dyson}), (\ref{eq:sigma}) and
370: (\ref{eq:bethesalpetereh}) are schematically represented in
371: Fig. \ref{fig:diagrams} for both the SCBA and the SCLA.
372: \begin{figure}
373: \centerline{\psfig{figure=ptfig2.eps,width=8truecm}}
374: \caption{(a) The Dyson equation, thick lines are the dressed
375: one-particle Green function, thin lines the bare ones. (b) The
376: self-energy with the scattering kernel $T(1;2)$, for the on-site
377: Coulomb potential. (c) The scattering kernel $T(1;2)$ for the Born
378: approximations. (d) The scattering kernel $T(1;2)$ in the ladder
379: approximation, given by the Bethe-Salpeter equation.}
380: \label{fig:diagrams}
381: \end{figure}
382: 
383: 
384: In the self-consistent approximation we solve self-consistently
385: Eqs.(\ref{eq:dyson}) and (\ref{eq:sigma}) with the scattering kernels
386: given by Eq. (\ref{eq:tborn}) or (\ref{eq:bethesalpetereh}) in the
387: Born or SCLA case respectively.  We explain in details how numerical
388: self consistency for a fixed temperature and chemical potential $\mu$
389: has been implemented.
390: 
391: 
392: (i) We start with ${G^+}(k,\omega)$ of the form ${G^+}_{0}=(\omega-
393: k^2/4+i \gamma)^{-1}$, and ${G^<}_0(k,\omega)=-2 \Im
394: [{G^+}_{0}(k,\omega)]~f(\omega-\mu)$.
395: 
396: (ii) In order to take advantage of the local character of the
397: electron-hole interaction, we use the Fourier transform of the Green
398: functions of Eq. (\ref{eq:greens}), $G(1,2)$, which actually have only
399: a spatial dependence on $r=x_1-x_2$ as the system is homogeneous.  We
400: calculate the free pair propagator defined as
401: \begin{equation}\label{eq:hr}
402:  H(1;2)= i G(1,2) G(1,2)
403: \end{equation} 
404: The retarded and correlation parts are calculated passing from the
405: Keldysh contours to the usual time axis:
406: \begin{equation}\label{eq:hr<}
407:  H^<(r,t)= i G^<(r,t) G^<(r,t),
408: \end{equation}\begin{equation}
409:  H^+(r,t)= i G^+(r,t) G^+(r,t)+
410: \label{eq:hr+} 2 i G^+(r,t) G^<(r,t).
411: \end{equation}
412: The dependence on the relative distance has been explicitly
413: shown. The dependence on the relative propagation time $t=t_2-t_1$
414: stems from the stationary condition. The pair propagator is then
415: Fourier transformed to $(k,w)$ space.
416: 
417: (iii.a) For the Born approximations, we use Eq. (\ref{eq:tborn})
418: \begin{mathletters}
419: \begin{eqnarray}
420: T^+(q,\omega)=-1+H^+(q,\omega)\\ T^<(q,\omega)=H^<(q,\omega)
421: \end{eqnarray}
422:  \end{mathletters}
423: 
424: (iii.b) For the SCLA case, Eq. (\ref{eq:bethesalpetereh2}) is readily
425: solved as:
426: \begin{mathletters}
427: \begin{eqnarray}\label{eq:t+w}
428: T^+(q,\omega)=-(1+H^+(q,\omega))^{-1}~,\\
429: T^<(q,\omega)=|T^+(q,\omega)|^2 H^<(q,\omega)~.
430: \end{eqnarray}
431: \end{mathletters}
432: 
433: (iv) The electron self-energy is then calculated in real space as in
434: Eq. (\ref{eq:sigma}). The retarded and correlation functions, defined
435: on the usual time axis, read:
436: $$
437: \Sigma^<(r,t)= i T^<(r,t)G^>(-r,-t)=-i T^<(r,t)[G^>(r,t)]^*,$$
438: $$
439: \Sigma^+(r,t)= - i T^<(r,t) G^-(-r,-t) +i T^+(r,t) G^<(-r,-t)=
440: $$
441: $$= - i T^<(r,t) [G^+(r,t)]^*+ i T^+(r,t) [G^<(r,t)]^*.$$
442: 
443: (v) The new electron Green's functions are then calculated as
444: \begin{mathletters}
445: $$
446: G^+(k,\omega)=(\omega- k^2/4+\Sigma^+(k,\omega)+i\gamma)^{-1},$$$$
447: G^<(k,\omega)=-2 \Im [G^+(k,\omega)]~f(\omega-\mu).$$
448: \end{mathletters} 
449: 
450: 
451: 
452: The procedure is repeated through step (ii) until self-consistence is
453: reached. Fast Fourier transforms are performed on a finite grid of
454: 16384$\times$128 points for the frequency and wavevector domains
455: respectively.  The frequency and wavevector domain $(k,w)$ is
456: (-20,20)X(-3,3).  Due to the finite k-range, we obtain $E_b=0.4$. The
457: external $\gamma>0$ in (v) is adiabatically switched off during 
458: self consistency. In this way the imaginary part of the final $G$ is
459: provided by the interaction only. For the non-self-consistent Born
460: approximation (or Markov-Born), we use the frequency independent,
461: on-pole $\Sigma^+_{k}(\omega=k^2/4)$ in the Green function, and stop
462: the procedure at step (v).
463: 
464: All of the considered approximations are conserving in the sense of
465: Kadanoff and Baym:\cite{baym61} the total density, total momentum,
466: total energy and total angular momentum of the system are
467: conserved. This is a fundamental property of any approximation for the
468: self-energy, otherwise unphysical results may result.  We
469: will come back to this point again in the calculation of the optical
470: response of the system in the various approximations.
471: 
472:   The polarized electron-hole gas with a contact potential interaction
473: that we consider here can be mapped onto a single band Hubbard model
474: with spin 1/2 and attractive interaction, in the limit of infinite
475: width of the band, and infinite on-site interaction so as to produce
476: finite $E_b$ (continuum limit). For the Hubbard model with repulsive
477: interaction, the self-consistent ladder approximation is known as
478: fluctuation exchange approximation (FLEX).\cite{flex} Even though in
479: this repulsive case the opening of an Hubbard gap is not reproduced,
480: in the attractive case the gap is of different nature and is
481: notoriously well described within this approximation.\cite{kagan98}
482: 
483: 
484: 
485: \section{Single Particle properties}
486: \label{sec:single}
487: 
488: \begin{figure} 
489: \centerline{\psfig{file=ptfig3.eps,width= 9 cm}}  
490: \caption{Electron spectral function at $k=0$ as a function of the
491: total carrier density. $T=0.1$.}
492: \label{fig:spectral1}
493: \end{figure}
494: We plot in Fig. \ref{fig:spectral1} the electron spectral function at
495: $k=0$ for different densities, obtained in the SCLA. The temperature
496: $T=0.1\ll E_b=0.5$. For this particular figure we used the k-range
497: (-6,6) in the numerical solution. For $n\sim 0.001$ the electron
498: spectral function has a very narrow Lorentzian shape.  For larger
499: densities, a satellite structure appears in the low energy side of the
500: spectral function. This structure is located at the exciton binding
501: energy, below the main peak, and accounts for the correlation of the
502: electron at $k_e=0$ bound with holes in other $k_h$ states. Excitons
503: with all values of the center of mass wavevectors are involved in this
504: peak,\cite{zimmermann88} as we will also show in the next section.
505: This correlation structure is of course not present in the Born
506: approximations, which maintain their single quasi-particle peak
507: structure at any density. As the density rises, the relative weight of
508: the satellite structure in the SCLA spectral function with respect to
509: the main quasi-particle peak increases. Moreover, both structures
510: become broader. For densities $n>0.1$, it becomes difficult to
511: distinguish between the excitonic and the main peak, and a single
512: broad, red-shifted quasi-particle structure appears. The merging of
513: the excitonic peak with the single quasi-particle peak can be
514: interpreted as a Mott transition. We remark however that it is
515: difficult to give a rigorous estimate of the Mott density, because
516: broadening is gradual and a sharp threshold is not evident.
517: 
518: 
519: 
520: \begin{figure} 
521: \centerline{\psfig{file=ptfig4.eps,width=10 cm}}
522: \caption{Self energy $\Sigma (k=0,\omega) $ for $T=0.1$ and different
523: densities given in the Figure, in the self consistent Born
524: approximation (SCBA) and self-consistent ladder approximation
525: (SCLA).}
526: \label{fig:self}
527: \end{figure}
528: In Fig. \ref{fig:self} we show the imaginary part of the electron self
529: energy for different densities. We will also refer in the following to
530: this quantity as (energy dependent) broadening. In the SCLA
531: $\Im(\Sigma)$ shows two peaks corresponding to dephasing experienced
532: by an unbound or a bound state propagated in the system. Both peaks
533: increase with density. At low density, $n=0.01$, dephasing at the
534: exciton energy dominates, while at $n=0.05$ both peaks become
535: comparable.  In the SCBA, the peak of $\Im(\Sigma)$ at $-E_b$ is
536: obviously absent, as propagation of bound pairs is not allowed in the
537: theory. Instead at low density $n=0.01$, the broadening at the main
538: quasi-particle energy ($\omega=0$) is similar in the SCLA and
539: SCBA. Indeed, few excitons are expected in the plasma, and scattering
540: mainly originates from free carriers.  An heuristic understanding may
541: be obtained with the action mass law relating the concentration of
542: different chemical species in a reaction. In our case, electron+hole
543: $\leftrightarrow$ exciton, and the law states ${n_c}^2/n_X=n^*(T)$,
544: where $n_c$ and $n_X$ represent the density of unbound carriers and
545: excitons respectively. $n^{*}(T)$ is a cross-over density which
546: depends only on the temperature, and at T=0.1, $n^{*}\sim 0.005$ . For
547: $n < n^*$, $n_X \ll n$. Therefore, a small density of excitons is
548: expected at $n=0.01\sim n^*$. At this density the exciton-free carrier
549: contribution to free carrier broadening is weaker than the free
550: carrier-free carrier contribution.  At larger density, bound excitons
551: in the plasma become dominant, and broadening at the quasi-particle
552: peak becomes much larger in the SCLA than in SCBA. This trend is
553: observed up to $n=0.15$ indicating relevance of correlated states in
554: the plasma even at this large density. However, the heuristic
555: interpretation of the broadening based on the action mass law given
556: above becomes meaningless, as it is not possible to distinguish
557: between two chemical species anymore.  At the highest considered
558: density $n=0.4$, we clearly observe a dip to very small broadening in
559: the self-energy for SCBA case. This dip is exactly at the Fermi energy
560: and accounts for the blocking of the scattering at the Fermi level. In
561: fact, only at this large density the Fermi gas becomes degenerate
562: (Fermi energy much larger than T). In the SCLA, the broadening never
563: vanishes, but is only partially reduced at the Fermi level. This is
564: indicative of a more complicated structure of the electron-hole plasma
565: and of the broadening process, where Pauli blocking looses much of its
566: effectiveness, and is suggestive of a non-Fermi liquid behavior.  We
567: finally noticed that $\Re(\Sigma)$ is comparable to $\Im(\Sigma)$ at
568: any density in the SCLA, and thus remarked that the excitonic
569: satellite peak in the spectral function does not correspond to a zero
570: of $\omega-\Re(\Sigma(\omega))$. Indeed, we do not expect appearance
571: of another simple quasi-particle in the plasma. For $\omega\sim -E_b$,
572: and for $E_b>>|\Im(\Sigma)|$, we have $ A_{k=0}\sim
573: |\Im(\Sigma)|/|E_b|^2$. Thus, the correlation satellite in the
574: spectral function follows the peak of $\Im(\Sigma)$ at $\omega=-E_b$.
575: 
576: 
577: In Fig.\ref{fig:mucomp} we show the density dependence of the electron
578: (or hole) chemical potential as a function of the density for $T=0.1$,
579: in the three considered approximations. At very low density, $n \ll
580: 0.01$, the chemical potential is $ \mu \ll - 0.4 $, and its value is
581: similar in all approximations. In this case, we are describing free
582: electron-hole pairs, as also suggested by a law of mass action, which
583: gives a crossover density of about $n^*=0.005$ at T=0.1. In this case
584: the description of $\mu$ is rather good even in approximations that
585: neglect existence of bound states. 
586: \begin{figure} 
587: \centerline{\psfig{file=ptfig5.eps,width= 8 cm}}  
588: \caption{Chemical potential $\mu$ in the three approximations, at
589: T=0.1.  The short-dashed line shows half the chemical potential of
590: non-interacting bosons, with ground state energy of $-E_b$.  }
591: \label{fig:mucomp}
592: \end{figure} 
593: However, for $n > 0.01$, the
594: chemical potential in the Born approximations is much larger than for
595: the SCLA. In fact, in the SCLA we are also describing the fraction of
596: cold interacting excitons in the plasma.  As the ground state energy
597: of this part of the plasma is at about $-E_b$, the total energy of the
598: system is thus reduced with respect to that of a gas of unbound
599: particles. In order to clarify further this point, we plot on the same
600: graph half of the chemical potential of bosons at a ground state
601: energy of $-E_b$ and mass equal to the exciton mass. 
602: %(2 in our units). 
603: This is the chemical potential of a gas of electron (and
604: holes) completely bound into bosonic excitons. The halving comes from
605: the equilibrium condition $\mu_{X}=\mu_e+\mu_h=2 \mu$.  The chemical
606: potential of bosons compares reasonably well with the chemical
607: potential calculated in the SCLA, up to $\mu<-0.25$, $n<0.1$. Above
608: this limit, the chemical potential from the SCLA grows faster. There
609: are two reasons for this faster growth: first, the exciton gas is
610: repulsively interacting, second, the exciton gas eventually undergoes
611: a Mott transition and ionizes into free carriers. At densities
612: $n>0.2$, the SCLA and SCBA are indeed comparable, and at even larger
613: density, also the MBA approximation is reasonable. This shows that the
614: SCLA is a powerful tool for the investigation of the intermediate
615: regime of densities, where deviations from both the (non-interacting)
616: bosonic and pure fermionic models are important.
617: 
618: 
619: 
620: 
621: 
622: \section{Mapping the SCLA to a bosonic model at low density and temperature}
623: \label{sec:excpot}
624: 
625: 
626: 
627: In Fig.\ref{fig:spectral2} we plot $A(k,\omega)$ for $T$=0.043 and
628: n=0.02. We can observe the parabolic dispersion of the main
629: quasi-particle peak, broadened by the scattering (white region in the
630: plot). Broadening is larger at small $k$ due to the larger phase space
631: in one dimension. For $\omega<0$, we have the correlation structure
632: shown for $k=0$ in Fig. \ref{fig:spectral1}, which is also present at
633: larger $k$.  We remark that this region of correlated electrons
634: extends into a region of $k$ of the order of 1 (i.e. of
635: $a_B^{-1}$). Its shape (dispersion) is related to the exciton
636: wavefunction, as we show in the following.
637: 
638: \begin{figure} 
639: \centerline{\psfig{file=ptfig6.eps,width= 8. cm}}
640: \caption{Contour plot of $A_e(k,\omega)$ at $T=0.04$ and n=0.02.}
641: \label{fig:spectral2}
642: \end{figure}
643: 
644: In a low-density, low-temperature limit, we can start to consider
645: $G=G_0$, and analytically calculate the retarded pair propagator
646: $H^+(k,\omega)$ from Eq. (\ref{eq:hr+}), neglecting $G_0^<$, which is
647: of the order of the density. We obtain:
648: $$ H^{(0)+}(k,\omega)\sim
649: \frac{-1}{\sqrt{2}\sqrt{\omega-\frac{k^2}{8}+2 i \gamma}},$$ where
650: $\gamma>0$ is the usual regularization number. Solving for the
651: T-matrix with the Bethe-Salpeter equation (\ref{eq:t+w}), and
652: expanding around its pole we obtain:
653: $$T^{(0)+}(k,\omega)=\frac{1}{\omega+\frac{1}{2}-\frac{k^2}{8}+2 i
654: \gamma}.$$ Thus, the T-matrix at the lowest order for the contact
655: potential has the form of a free propagator, for particles at an
656: energy $-1/2+k^2/8$, having a mass which is twice the electron mass,
657: i.e. the electron plus hole mass. Expanding the Bethe-Salpeter
658: equation to higher order, we obtain:
659: $$T^{(1)+}(k,\omega)=T^{(0)+}(k,\omega) H^{(1)+}(k,\omega)
660: T^{(0)+}(k,\omega).$$ Therefore, we may interpret $T^{(0)+}(k,\omega)$
661: as the free exciton (boson) Green function, and $ H^{(1)+}(k,\omega)$
662: as the lowest order self-energy.\cite{haussman} In particular
663: \begin{equation}\label{eq:hr1}
664: H^{(1)}(1,2)=2 i G^{(1)}(1,2) G_0(1,2),\end{equation}
665: $$G^{(1)}(k,\omega)= G_0(k,\omega) \Sigma^{(0)}(k,\omega)
666: G_0(k,\omega),$$
667: \begin{equation}\label{eq:sigmar0}
668: \Sigma^{(0)}(1,2)= i T^{(0)}(1,2) G_0(2,1).\end{equation} As
669: $T^{(0)}(k,\omega)$ is peaked close to $\omega=-1/2$, we may neglect
670: $G_0^<(k,\omega)$ in this region, and use
671: $$\Sigma^{(0)+}(k,\omega)\sim i\int\frac{dqd\omega'}{(2\pi)^2}
672: T^{(0)<}(q,\omega') G_0^{+}(k+q,\omega+\omega').$$ Integrating around
673: $\omega'=-1/2$, we obtain:
674: $$\Sigma^{(0)+}_k(\omega)\sim
675: \frac{n}{\omega+\frac{1}{2}+\frac{k^2}{8}+i\gamma},$$ where $n$ is the
676: total density. This is the structure shown in Fig. \ref{fig:self} at
677: $\omega=-1/2$, which produces the satellite correlation peak in the
678: electron spectral function, shown in Fig. \ref{fig:spectral2}. We also
679: notice that this structure has {\em negative} dispersion, as also
680: apparent in Fig. \ref{fig:spectral2}.
681: 
682: \begin{figure}
683: \centerline{\psfig{figure=ptfig7.eps,width=8truecm}}
684: \caption{The perturbation expansion to lowest order for the boson
685: (exciton) self-energy, and the four-particle vertex $F^{(0)}$.  Thin
686: lines are the bare electron (or hole) propagator. The shaded
687: two-particle Green function is the bare exciton propagator.}
688: \label{fig:excexc}\end{figure}
689: 
690: We can now define an exciton-exciton interaction at low density, from
691: the Eqs. (\ref{eq:hr1}-\ref{eq:sigmar0}) above, which are pictorially
692: shown in Fig. \ref{fig:excexc}. We thus define a four-point
693: interaction kernel $F^{(0)}$, and the boson self-energy $H^{(1)}(1,2)$
694: is written as:
695: \begin{equation}
696: H^{(1)}(1,2)= \int d\bar{1}~ d\bar{2}~F^{(0)}(1,2;\bar{1},\bar{2})
697: T^{(0)}(\bar{2},\bar{1}). \end{equation} We can see from
698: Fig. \ref{fig:excexc} that $F^{(0)}$ represents the electron (or hole
699: exchange) in the scattering of the two excitons. Thus
700: \begin{equation}
701:  F^{(0)}(1,2;1'2') = 2 i G_0(1,2) G_0(1',2') G_0(1,2') G_0(1',2).
702: \end{equation} If we neglect retardation
703: (or memory) effects, we may define an instantaneous potential at the
704: exciton energy using $\tau_1=\tau_2$, $\tau_{1'}=\tau_{2'}$, and
705: calculating the resulting $F$ at the exciton frequency
706: $\omega_0=-1/2=E_B$. For simplicity, we also use $k,k'\sim 0$ for the
707: incoming excitons, and arrive to
708: $$F^{(0)}(q,\omega_0) \sim 2i \sum_{q',\omega'}
709: G_0^{+}(q',\frac{\omega_0}{2}+\omega')
710: G_0^{+}(-q',\frac{\omega_0}{2}-\omega')$$ \begin{equation}
711: G_0^{+}(q-q',\frac{\omega_0}{2}+\omega')
712: G_0^{+}(q'-q,\frac{\omega_0}{2}-\omega')=\end{equation}
713: $$=\frac{64}{q^2}\frac{2\sqrt{4+q^2}-(4-q^2)}{(4+q^2)^2}.$$ This
714: function rapidly drops to zero at $q>1$ as expected, and
715: $F^{(0)}(0,\omega_0)=6$.  This value is also obtained from standard
716: boson-boson exchange-interaction expressions (see
717: e.g. Ref. \onlinecite{haughanamura}) when only the electron-hole
718: interaction is used.  We show the full structure of
719: $F^{(0)}(q,\omega_0)$ In Fig. \ref{fig:vq}.
720: \begin{figure}
721: \centerline{\psfig{figure=ptfig8.eps,width=8truecm}}
722: \caption{ The exciton-exciton interaction potential $W_q=F^{(0)}(q,\omega_0)$.}
723: \label{fig:vq}
724: \end{figure} 
725: 
726: 
727: We are now in a position to discuss some of the qualitative changes
728: introduced by simplifying the Coulomb interaction to an on-site
729: one. In the limit of low density, we may in fact compare the
730: exciton-exciton interaction calculated above with that calculated
731: using a realistic Coulomb interaction, as a standard boson-boson
732: exchange expression at small momenta exists.\cite{tassone99xx} In a
733: realistic wire, a short-range cut-off is introduced by the finite size
734: of the electron and hole wavefunctions in the confinement directions.
735: A typical cut-off is of the order of the Bohr radius or
736: smaller. Typically, we have to consider tight confinements in order to
737: avoid participation of higher confined levels into the exciton
738: wavefunction. The Coulomb interaction is then reasonably represented
739: by a function $V_r=A/(r+r_0)$, where $r_0$ is a cut-off distance, and
740: $A$ is then chosen such that $E_b=-1/2$. 
741: %Using $r_0=1$, we have
742: %$A=0.65$ and a resulting Bohr radius of $0.4$ only. In this limit, the
743: %electron and hole wavefunctions are not tightly confined.
744: We used $r_0=0.1$, and $A=0.18$. In this case the resulting Bohr radius
745: is 1, indicating tight confinement of the carriers in the confinement
746: plane. The exciton-exciton exchange interaction at small $q$ is
747: calculated in the small $q \ll {a_B}^{-1}$ limit
748: $$W_{xx}=2\sum_{k,k'}V_r(k-k')\phi_{1s}(k)\phi_{1s}(k')$$
749: \begin{equation}
750: \left[|\phi_{1s}(k)|^2-\phi_{1s}(k)\phi_{1s}(k')\right]\sim
751: 1.4~, \label{eq:xxbose}
752: \end{equation}
753: where $\phi_{1s}(k')$ is the 1s wave-function of the exciton. The
754: positive term in Eq. \ref{eq:xxbose} is the exchange term due to the
755: attractive electron-hole interaction, while the negative term is due
756: to the electron-electron and hole-hole repulsive interactions.  In the
757: delta like-potential the negative term is absent because of
758: locality. In the long range case instead, the electron-electron (and
759: hole-hole) interaction largely cancel the electron-hole interaction in
760: the exchange integral, resulting in an exciton-exciton interaction
761: which is four times smaller when compared with the value of 6 obtained
762: for $F^{(0)}(q=0,\omega_0)$. We conclude that local interaction leads
763: to an over-estimation of the boson-boson interaction at a given
764: density, and therefore of broadening. However, this does not imply
765: that broadening effects are {\em qualitatively} different. Moreover,
766: we also stress that this overestimation does not concern at all the
767: comparison of the SCLA with the other simpler approximations
768: considered in this work, as all are carried out using the same
769: interaction potential.
770: 
771: 
772: 
773: \section{Optical Properties}
774: \label{sec:optical}
775: 
776: We now consider interaction of the electron-hole system with a
777: transverse electromagnetic field in the dipole approximation:
778: \begin{equation}
779: H_P=\sum_{{\bf q },q} C_{{\bf q},q} (a_{{\bf
780: q},q}P^{\dagger}_{q}+a^{\dagger}_{{\bf q},q}P_{q}),
781: \label{eq:Hp}
782: \end{equation}
783: where
784: $$ C_{{\bf q},q}=\frac{e p_{cv}}{m_0 c}\sqrt{\frac{4\pi \hbar c}{ q_t
785: V}} I_{\parallel}({\bf q}),
786: $$
787: is the dipole matrix element, which includes also the overlap integral
788: of the electromagnetic field with the confined carriers
789: $I_{\parallel}$.  Here ${\bf q }$ represents 2D wavevectors in the
790: plane orthogonal to the wire axis, $q$ are wavevectors along the wire,
791: while $q_t^2={\bf q}^2+q^2$, $m_0$ is the free electron mass, $c$ the
792: velocity of light, and $p_{cv}$ the momentum matrix element between
793: the conduction and hole bands.  $a_{{\bf q},q}$ is the photon
794: destruction operator, and $P_{q}$ is the polarization operator,
795: defined as $P_{q}=\sum_{k}d_{q-k}c_{k}$, i.e. local in real space in
796: the dipole approximation.  As the section of the wires is always much
797: smaller than the wavelength of light, we have $I_\parallel\sim
798: 1$. Moreover $q_t \sim E_{GAP}/(\hbar c)$ is almost constant, thus we
799: use arbitrary units from now on, with $C=1$. Actual values of
800: absorption and gain are easily calculated by using the appropriate
801: value of $C$ for the given material.
802: 
803: 
804: 
805: In the process of scattering of light from the system, an absorbed
806: photon creates a coherent electron-hole pair which propagates in the
807: crystal, interacts with the background plasma, then recombines, and a
808: photon is emitted.  When we introduce photon propagators, the
809: interaction process is represented by a photon self-energy. We will
810: not solve the photon Dyson equation, nor we will dress the electron
811: and hole propagation with the photon interaction. In other words, we
812: neglect both polaritonic effects, and higher order non-linear
813: interactions. In fact, we are restricting our attention to weak
814: external fields, i.e. assuming interaction of the plasma with the
815: photon field to be weak. Otherwise, it could be possible for the
816: system to be driven out of equilibrium, because the photon chemical
817: potential is zero and far below that of the elecron-hole plasma. This
818: equilibrium condition is well fulfilled in real systems, exception
819: made for the strongest laser fields.  Within this weak interaction
820: approximation, the photon self-energy is exactly given by the pair
821: correlation function
822: \begin{equation}
823: \Pi_q(\tau,\tau')= \langle T P^{\dagger}_{q}(\tau)
824: P_{q}(\tau')\rangle.\label{eq:pairprop}\end{equation} However, this
825: correlation function has to be calculated within some approximation,
826: and also using approximate propagators.  The resulting self-energy may
827: not be conserving. In particular, particle flux conservation at the
828: vertex translates in the longitudinal f-sum rule\cite{baym61}, which
829: has to be respected in any reasonable approximation.
830: 
831: In order to shed light on this important issue, we consider the
832: problem of interaction of photons with the plasma from a different
833: viewpoint. As we are only interested in the linear response to the
834: external electromagnetic field, the correct photon self-energy is also
835: given by the linear response of the plasma to a {\em classical}
836: electromagnetic field, i.e. considering the photon creation and
837: destruction operators in Eq. (\ref{eq:Hp}) as c-numbers.  Baym and
838: Kadanoff give a recipe to construct a conserving expression for this
839: response function in Ref.\onlinecite{baym61}. When an external
840: coherent electromagnetic field is applied to the plasma, coherence
841: develops in the plasma, and pair propagators have additionally to be
842: defined along with the single particle propagators. They are analogous
843: to anomalous propagators in the theory of superconductivity. The
844: electron-hole anomalous propagator reads
845: \begin{equation}\label{eq:pairprop2}
846: G_{eh}(1,2)=i\langle T d(2) c(1) \rangle. \end{equation} 
847: Times are defined on the Keldysh contour as usual. Another anomalous
848: propagator $G_{he}$ can be similarly defined for convenience, and
849: notation regrouped into a matrix one, like for Nambu propagators in
850: the problem of superconductivity.\cite{nambu60} Dyson
851: Eq. (\ref{eq:dyson}) is extended to include both 
852: anomalous propagators, and the external interaction $H_p$:
853: $$
854: \int d\bar{2}~G_0^{-1}(1,\bar{2})G(\bar{2},3)=\delta(1,3)+\int
855: d\bar{4}~\Sigma(1,\bar{4})G(\bar{4},3)+$$
856: \begin{equation}
857: +\int
858: d\bar{4}~\Sigma_{eh}(1,\bar{4})G_{he}(\bar{4},3)-A(1)G_{he}(1,3)\end{equation}
859: $$\int d\bar{2}~G_0^{-1}(1,\bar{2})G_{eh}(\bar{2},3)=\int
860: d\bar{4}~\Sigma(1,\bar{4})G_{eh}(\bar{4},3)+$$
861: \begin{equation}+\int d\bar{4}~\Sigma_{eh}(1,\bar{4})G(\bar{4},3)- A(1)G(3,1)
862: \label{eq:dysonanomal}\end{equation}
863: 
864: As a general rule for generating a conserving approximation, the self
865: energies have to be obtained from functional derivation of a
866: grand-potential with respect to the propagators. The normal
867: self-energies have been already introduced.  For the anomalous self
868: energy $\Sigma_{eh}(1,2)$, we need to generalize the expression of the
869: grand-potential including terms where the normal propagators are
870: replaced by the anomalous ones.  The anomalous propagators $G_{eh}$
871: are at least first order in the external potential $A$, as spontaneous
872: symmetry breaking is ruled out by the Mermin-Wagner theorem.  For
873: linear response, we need the anomalous self energy $\Sigma_{eh}$ up to
874: linear terms in the external potential only, thus at most linear terms
875: in the anomalous propagators. Consequently, we need at most quadratic
876: terms in these propagators in the grand-potential to obtain the
877: $\Sigma_{eh}$ from the functional derivative. We conclude that the
878: only and sufficient term to be considered is the anomalous Fock term,
879: which is the usual Fock term with anomalous propagators replacing
880: normal ones. As the approximation is conserving, we are also
881: guaranteed that absorption fulfills the f-sum rule\cite{baym61} for
882: any value of density and temperature.  The anomalous self-energy
883: obtained from the anomalous Fock term reads
884: \begin{equation}\Sigma_{eh}(1,2)=-i G_{eh} (1,2)+O(A^2).\label{eq:sigmaanomal}
885: \end{equation}
886: Normal $G$ are at least second order in the external potential $A$,
887: while both $G_{eh}$ and $\Sigma_{eh}$ are at least first order. The
888: anomalous Dyson Eq. (\ref{eq:dysonanomal}) to lowest order becomes:
889: $$\int d\bar{2}~G^{-1}(1,\bar{2})G_{eh}(\bar{2},3)=-i \int
890: d\bar{4}~G_{eh}(1,\bar{4})G(\bar{4},3)-$$
891: \begin{equation}
892: - A(1) G(1,3).
893: \label{eq:dysonanomal2}\end{equation}
894: Introducing the response function
895: \begin{equation}
896: \tilde \Pi (1,2)= -i \left.\frac{\delta
897: G_{eh}(1,1)}{A(2)}\right|_{A=0},
898: \end{equation}
899: and taking the first order variation of Eq. (\ref{eq:dysonanomal2}),
900: we obtain and equation for $\tilde \Pi (1,2)$:
901: $$
902: \tilde \Pi (1,2)= i G(1,2)G(1,2) -i\int d\bar{3}~ G(1,\bar{3})
903: G(1,\bar{3}) \tilde \Pi (\bar{3},2)=$$
904: \begin{equation}=
905: H(1;2) - \int d\bar{3}~ H(1;\bar{3}) \tilde \Pi (\bar{3},2),
906: \end{equation}
907: which is easily solved and gives
908: \begin{equation} 
909: \tilde \Pi (1,2)= H(1;2) +\int d\bar{3}~ d\bar{4}~ H(1;\bar{3}) T
910: (\bar{3};\bar{4}) H (\bar{4};2). \end{equation} It is straightforward
911: to show that $\tilde \Pi (1,2)$ obeys bosonic Kubo-Martin-Schwinger
912: relations, as both $H$ and $T$ do. In particular, we have:
913: \begin{eqnarray} 
914: \tilde \Pi^+(q,\omega)= H^+(q,\omega)\left|  T^+(q,\omega)\right|^2 \\ 
915: \tilde \Pi^<(q,\omega)= H^<(q,\omega)\left|  T^+(q,\omega)\right|^2.
916: \end{eqnarray}
917: Absorption $\alpha(\omega)$ and photoluminescence $PL(\omega)$ at
918: $q=0$ are derived as $\alpha(\omega)=-\Im [\Pi^+(0,\omega)]$ and
919: $PL(\omega)=-\Im [\Pi^<(0,\omega)]$ respectively.
920: 
921: We remark that higher order terms than the anomalous Fock one have to
922: be considered in the anomalous self-energy expansion when a {\em
923: finite} external field is present, such as in the case of a strong
924: laser field.  Anomalous Born terms calculated in the Markov
925: approximation have been considered in the
926: semiconductor Bloch equations by Lindberg and Koch in
927: Ref. \onlinecite{lindberg}, and named polarization-polarization
928: scattering terms. Here we do not further pursue such extensions, which
929: are clearly beyond the scope of the present paper. We only remark that
930: our approach is easily extended to such conditions, while keeping full
931: control of the conservation laws.
932: 
933: 
934: \subsection{Optical spectra and stability of the excitonic emission}
935: 
936: We show in Fig. \ref{fig:abs}, the absorption spectra in the normal
937: direction ($q=0$) at $T=0.2$, for different densities, in the MBA,
938: SCBA, and SCLA. In the absorption spectrum, we remark at low density
939: the characteristic excitonic peak and the continuum absorption. The
940: exciton linewidth is extremely narrow in the SCBA, and much broader in
941: the MBA. The SCLA linewidth is intermediate. In the SCBA there is a
942: very small broadening at the exciton energy, as no excitons are
943: represented in the theory, while in the MBA, the free carrier
944: broadening at $\omega=0$ is assumed. This is somewhat larger than the
945: broadening in the SCLA at $T=0.2$.  For large density $ n>0.2$, the
946: excitonic peak is bleached in all approximations, and a region of
947: negative absorption, i.e. optical gain, appears. The absorption
948: changes sign at $\omega=\mu_e+\mu_h=2\mu$, coinciding with the change
949: of sign of the Bose factor, and resulting in a well defined
950: -i.e. positive- photoluminescence through the Kubo-Martin-Schwinger
951: relation.
952: 
953: We plot the emission or PL spectra in Fig. \ref{fig:pl}. First we
954: notice the different scale for the MBA, where a smaller peak emission
955: is calculated even for the largest considered density.  For large
956: density $n>0.2$, we observe a saturation of the intensity. This can be
957: understood in terms of the fermionic nature of the carriers, when the
958: electron-hole plasma becomes degenerate. Second, we notice
959: exciton-like emission even when excitonic absorption is bleached in
960: all models.  In fact, vertex corrections place the free carrier
961: emission at the exciton energy, even when no bound excitons are
962: described in the gas. 
963: \begin{figure}
964: \centerline{\psfig{file=ptfig9.eps,width=11 truecm}}
965: \caption{Absorption spectra in the Markof-Born Approximation (MBA),
966: self consistent Born approximation (SCBA) and self consistent ladder
967: approimation (SCLA) for different densities indicated in the figure.}
968: \label{fig:abs}
969: \end{figure}
970: Eventually, only at very small density the free
971: carrier and excitonic emission become distinguishable, as we notice an
972: exponential emission shoulder at high energy $\omega>0$ in the SCLA
973: and SCBA, reminding of a fermionic emission tail at small degeneracy
974: (Boltzmann distribution). At larger density, the two features at
975: different energy merge into a unique one, due to increased
976: broadening. A single peak is also observed in experiments, where
977: finite noise in the data and large inhomogeneous broadening mask any
978: minor feature.\cite{optwires} For this reason, it is in practice very
979: delicate to establish the position of the band-gap (bottom of the
980: free-carrier bands) directly from emission data.\cite{piermarocchi99}
981: Third, we remark that the low-energy shoulder of the PL emission in
982: the MBA is Lorentzian because of the Markov approximation, and the
983: broadening largely over-estimated also at low energies. The linewidth
984: in the SCBA is instead underestimated at low-density, as for
985: absorption.
986: 
987: \begin{figure} 
988: \centerline{\psfig{file=ptfig10.eps,width=11 cm}}
989: \caption{Photoluminescence spectra in the MBA, SCBA and SCLA for different
990: densities indicated in the figure.}
991: \label{fig:pl}
992: \end{figure}
993: 
994: 
995:  An interesting -and directly observed- physical quantity, is the
996: energy of the PL peak. As remarked above, emission is excitonic due to
997: the correlation of the coherent electron-hole pair emitted. In
998: Fig. \ref{fig:stability} we plot the position of this peak in the PL
999: as a function of the carrier density, together with the band-gap
1000: renormalization (BGR) defined as twice the energy of the main peak of
1001: the spectral function. Interestingly, exciton bleaching in absorption
1002: appears at a density which is comparable to that where the band gap
1003: renormalization crosses the emission energy, i.e. $n\sim 0.15$. This
1004: can therefore be assigned as a Mott density, i.e. the energy where the
1005: binding energy becomes negligible with respect to broadening, thus
1006: effectively vanishing. There are two interesting features to be
1007: observed in Fig. \ref{fig:stability}: first, the band gap
1008: renormalization is negligible in the SCLA for $n<0.1$, second,
1009: emission energy is constant in this range of densities, and
1010: blue-shifting less than in the SCBA for larger density. Comparison
1011: with the MBA is vitiated by excessive broadening in this
1012: approximation. The simple Hartree-Fock approximation instead gives
1013: results which are similar to those obtained in the SCBA for the band
1014: gap renormalization and the PL emission peak, apart from more
1015: pronounced blue-shifts.\cite{tassone99} Stability of the emission peak
1016: is usually interpreted as a partial compensation of the self-energy
1017: and vertex corrections. As in the Hartree-Fock approximation the
1018: broadening effects are missing, we deduce that these effects are
1019: indeed relevant for the cancellation found in the SCBA and in the SCLA
1020: at high density, and explain the reduction of the blue-shift with
1021: respect to simple Hartree-Fock calculations.
1022: \begin{figure} 
1023: \centerline{\psfig{file=ptfig11.eps,width=8 cm}}
1024: \caption{The band gap renormalization (BGR) and emission peak energy
1025: in the SCBA and SCLA as indicated in the figure.}
1026: \label{fig:stability}
1027: \end{figure}
1028: 
1029: 
1030: 
1031: \subsection{Excitonic gain}
1032: 
1033: 
1034: Optical gain is usually considered in the regime of degenerate
1035: electron-hole plasma. In this regime, spectral functions can be
1036: assumed to be simple Lorentzians, and the absorption can be written
1037: as:
1038: \begin{equation}\label{eq:gainfc}
1039: \alpha(\omega)\propto\int \frac{dk}{2 \pi} \frac{1-2 f(k^2/4
1040: -\mu)}{(\omega-k^2/2)^2+ \gamma^2},
1041: \end{equation}
1042: where $\gamma$ is the spectral broadening. Then, negative absorption
1043: or gain occurs only when the Pauli blocking factor ($1-2 f(k^2/4
1044: -\mu)$) becomes negative, which necessarily requires a chemical
1045: potential above the band gap, or in other words, inversion. However,
1046: both vertex corrections, and deviations of the spectral function from
1047: simple Lorentzian should be taken into account at lower density and
1048: temperature. In this case,
1049: absorption can be written as:
1050: \begin{eqnarray}
1051: \alpha(\omega) \propto |T^{+}(q=0,\omega)|^2
1052: \int\frac{dk}{2\pi}\frac{d\omega^\prime}{2 \pi}\times
1053: \label{eq:abstherm}
1054: \end{eqnarray}$$
1055: \left[1-f(\omega-\omega^\prime-\mu)-f(\omega^\prime-\mu)\right]
1056: A_{k}(\omega-\omega^\prime)A_{k}(\omega^\prime)~,
1057: $$
1058: The term in square brackets
1059: $1-f(\omega-\omega^\prime-\mu)-f(\omega^\prime-\mu)$ is the
1060: generalization of the Pauli blocking factor for a system where the
1061: quasiparticles are described by arbitrary spectral functions, and
1062: clearly one of the Fermi functions refers to electron occupation, the
1063: other to hole occupation. This expression of the absorption is valid
1064: only in the case of short range potential, and for the long range case
1065: $T^+$ depends also on the relative momentum of the pair $k$, and
1066: becomes part of the integral kernel.  Using the definition of Fermi
1067: and Bose functions, the absorption can be recast in the form
1068: \begin{eqnarray}
1069: \alpha(\omega)=\frac{|T^{+}(q=0,\omega)|^2}{g(\omega-2 \mu)}
1070: \int\frac{dk}{2\pi}\frac{d\omega^\prime}{2\pi} \times\label{eq:absg}
1071: \end{eqnarray}$$\times
1072: f(\omega-\omega^\prime-\mu)f(\omega^\prime-\mu)
1073: A_{k}(\omega-\omega^\prime)A_{k}(\omega^\prime)~,$$ where
1074: $g(\omega-2\mu)$ is the Bose function.  Therefore, gain clearly occurs
1075: below twice the chemical potential $\mu$ (i.e. $\mu_e+\mu_h$), even
1076: when the chemical potential is {\em below} the band-gap. However, for
1077: it to be sizeable, the spectral functions must have non-negligible
1078: weight in this region of the spectrum.  In the MBA and SCBA case, this
1079: weight is clearly given by broadening effects alone, while in the SCLA
1080: it is also due to the presence of the excitonic correlation peak in
1081: the spectral function as shown in Fig.\ref{fig:spectral1}.  A sizeable
1082: enhancement of the gain is found in correspondence to the excitonic
1083: resonance because of the vertex correction, given by the $|T^+|^2$
1084: factor in Eq. (\ref{eq:absg}). When the $2\mu< -E_B$, both gain and
1085: excitonic absorption coexist. In this case, we talk of {\it excitonic
1086: gain}. The system is inversion-less in the sense of
1087: Eq. (\ref{eq:gainfc}) above. However, only in the SCLA we are
1088: describing gain due to the presence and scattering of excitons in the
1089: low temperature, low-density plasma, while in the other
1090: approximations, gain is purely related to  dynamical effects in the
1091: interaction vertex with the photon.
1092: 
1093: In Fig.\ref{fig:gain} we plot the absorption spectra close to the
1094: exciton resonance at $T=0.1\ll E_B$. For the considered densities
1095: $n<0.1$, gain coexists with the excitonic resonance in absorption.
1096: Only for the MBA at $n=0.1$, the absorption peak is completely shifted
1097: to higher energies, and we can not anymore speak of excitonic
1098: absorption. As usual, the broadening in the exciton spectral region is
1099: too small and unphysical in the SCBA, resulting in very sharp and
1100: unphysical features. Gain in the SCLA is instead over a larger
1101: spectral region, and its value is larger than what predicted in the
1102: MBA at large density $n>0.3$ shown in Fig.\ref{fig:abs}. Excitonic
1103: gain clearly originates from inclusion of excitons in the plasma for
1104: the SCLA. In fact, reasonable broadening is calculated in the exciton
1105: spectral region, and mainly originating from exciton-exciton
1106: scattering as shown in Sec. \ref{sec:excpot}. We conclude, that the
1107: SCLA indicates that sizeable excitonic gain can be obtained at
1108: moderate density and temperature, in a reasonably large spectral
1109: region of the order of a fraction of $E_B$.
1110: 
1111: \begin{figure} 
1112: \centerline{\psfig{file=ptfig12.eps,width=11 cm}}
1113: \caption{Absorption spectra at $T=0.1$ for the MBA, SCBA, and SCLA,
1114: and low densities indicated in the figure.}
1115: \label{fig:gain}
1116: \end{figure} 
1117: Excitonic gain has been widely studied in II-VI quantum wells
1118: \cite{kozlov96}.  However these are two-dimensional system, and
1119: excitonic correlations are expected to be less pronounced in this
1120: systems, so at this stage it is not reasonable to make even a
1121: qualitative comparison.  For quantum wires, sizeable gain at 10 K
1122: ($T\sim 0.05$) for an estimated carrier density well below the Mott
1123: has been recently claimed by Sirigu {\it et al}.\cite{sirigu99} This
1124: is a first indication that excitonic gain might be relevant in these
1125: systems. However, a qualitative and quantitative understanding of this
1126: experiment is clearly premature. From the experimental side,
1127: inhomogeneous broadening due to interface disorder has to be decreased
1128: well below the binding energy, while a more realistic Coulomb
1129: interaction has to be addressed by the theory.
1130: 
1131: 
1132: 
1133: 
1134: \section{Conclusions}
1135: \label{sec:conclusions}
1136: 
1137: We have presented a model that includes excitonic correlation in the
1138: description of a highly-excited semiconductor quantum wire. The model
1139: has been simplified using a short range potential, and considering a
1140: polarized gas.  Correlation has been calculated self-consistently at
1141: the ladder level (SCLA). We have shown that bound states appear as a
1142: low-energy correlation peak in the spectral function of electrons and
1143: holes.  We compared the results obtained with SCLA to the ones
1144: obtained within lower order approximations that do not include
1145: excitonic correlation (Born approximations).  Even though the model is
1146: purely fermionic, we have shown how it can be effectively mapped at
1147: low temperatures and density to a gas of interacting excitons. We have
1148: thus derived an analytical expression for the effective
1149: exciton-exciton interaction.  The linear optical properties of the
1150: system have been calculated including vertex corrections at the Fock
1151: level. This ensures the conservation of sum rules in the optical
1152: response. The excitonic absorption at low density ensues both in
1153: ladder and Born approaches, but the broadenings at the exciton energy
1154: are either too small or too large in the Born approximations,
1155: depending on whether frequency dependence of the broadening is
1156: included or not. Excitonic emission well beyond exciton bleaching is
1157: also predicted in all models, but the peak shifts are more pronounced
1158: in the Born approximations, showing that a better cancellation between
1159: the self-energy and vertex correction results from the introduction of
1160: excitons. We have also shown that sizeable excitonic gain can be
1161: predicted at low temperature and density, when the electron plus hole
1162: chemical potential is just below the exciton energy. However, its
1163: correct description must include exciton broadening from
1164: exciton-exciton scattering, and unphysical values are thus obtained in
1165: the Born approximations. These qualitative conclusions clearly hold
1166: even for more refined descriptions of the electron-hole plasma than
1167: the Born approximations considered here, when these descriptions do
1168: not account for excitons in the plasma. For example this is the case
1169: for the model recently presented by Das Sarma and Wang in
1170: Ref. \onlinecite{dassarma00}, which includes screening at the
1171: plasmon-pole approximation level, but no excitons in the electron-hole
1172: gas. The excitonic gain calculated using this approximation shows a
1173: divergent behavior close to $-E_b$, similar to what we have found for
1174: the SCBA. The extension of the proposed approach to a full
1175: non-equilbrium theory could contribute to a deeper understanding of
1176: many body effects in electron-hole systems, also in the perspective of
1177: ultrafast optical experiments than can investigate far-from
1178: equilibrium conditions. In this context a recent theory by Hannewald
1179: {\it et al.}\cite{hannewald00} provides a convincing step in this
1180: direction, yet remaining within a non-dynamical approach.  The last
1181: issue concern the intrinsic limitations of the SCLA. It has been shown
1182: that in order to describe correctly the very diluite excitonic limit
1183: additional diagrams must be added to the self consistent ladder
1184: approach. An analysis of this problem can be found in
1185: Ref.\onlinecite{pieri98}, where a generalized T matrix approximation
1186: for the fermionic self energy overcoming this problem has ben
1187: proposed.  In conclusion, we have clearly shown that in the definite
1188: and important physical region of $T< E_b$ inclusion of excitonic
1189: correlation in the electron-hole plasma is relevant and necessary, and
1190: that the simplified model presented in this paper is a well-understood
1191: starting point for this purpose. Thus, its further developments to
1192: address realistic systems in higher dimensions are well motivated.
1193: 
1194: 
1195: \acknowledgments We thank A. Quattropani, P. Schwendimann, V. Savona,
1196: C. Ciuti, and L. J. Sham for stimulating discussions. One of the
1197: authors (C. P.) acknowledges support by the Swiss National Foundation
1198: for the Scientific Research.
1199: 
1200: 
1201: 
1202: 
1203: \begin{references}
1204: 
1205: \bibitem{rice77} Excellent reviews of the subject may be found in
1206: B. I. Halperin and T. M. Rice, in {\it Solid State Physics}, edited by
1207: H. Ehrenreich, F. Seitz, and D. Turnbull (Academic, Ney York, 1977),
1208: Vol.21.; T. M. Rice, {\it ibidem} Vol. 32.
1209: 
1210: \bibitem{haug76} H. Haug, Z. Phys. B {\bf 24}, 351 (1976).
1211: 
1212: \bibitem{zhu96} X. Zhu, M. S. Hybertsen, and P. B. Littlewood,
1213: Phys. Rev. B {\bf 54}, 13 575 (1996).
1214: 
1215: \bibitem{elliott57} R. J. Elliott, Phys. Rev. {\bf 108}, 1384 (1957).
1216: 
1217: \bibitem{haugkoch} H. Haug and S.W. Koch, 'Quantum theory of the
1218: optical and electronic properties of semiconductors, (Singapore, World
1219: Scientific, 1990).
1220: 
1221: 
1222: \bibitem{haug84} H. Haug and S. Schmitt-Rink, Progr. Quantum
1223: Electronics {\bf 9}, 3 (1984).
1224: 
1225: \bibitem{kraeft86} W. D. Kraeft, D. Kremp, W. Ebeling, and G. R\"opke,
1226: {\it Quantum statistics of charged particle systems}, Akademie-Verlag,
1227: Berlin 1986.
1228: 
1229: \bibitem{zimmermann88} R. Zimmermann, {\it Many Particle theory of
1230: highly excited semiconductors}, Teubner Liepzig (1988).
1231: 
1232: 
1233: \bibitem{pereira98} M. Pereira and K. Henneberger, Phys. Rev. B {\bf
1234: 58}, 1064 (1998).
1235: 
1236: 
1237: \bibitem{tassone99} F. Tassone and C. Piermarocchi,
1238: Phys. Rev. Lett. {\bf 82}, 843 (1999).
1239: 
1240: \bibitem{benner91} S. Benner and H. Haug, Euro Phys. Lett. {\bf 16},
1241: 570 (1991).
1242: 
1243: 
1244: \bibitem{tassone.93} F. Tassone and F. Bassani, Phys. Rev. B {\bf 51}
1245: 16973 (1995).
1246: 
1247: \bibitem{mahan81} D. G. Mahan, {\it Many particle physics}, Plenum
1248: Press (1981).
1249: 
1250: \bibitem{danielewicz} P. Danielewicz, Annals of Physics {\bf 152},
1251: 239, (1984).
1252: 
1253: 
1254: \bibitem{haug96} H. Haug and A. P. Jahuo, {\it Quantum kinetics in
1255: transpot and optical properties of semiconductors}, Springer-Verlag,
1256: Berlin (1996).
1257: 
1258: 
1259: \bibitem{kadanoff62} L. P. Kadanoff and G. Baym, {\it Quantum
1260: statistical mechanics}, W. A. Benjamin Reading, (1962).
1261: 
1262: 
1263: \bibitem{optwires} R. Cingolani {\it et al.} Phys. Rev. B {\bf 48}
1264: 14331 (1993); C. Gr\'eus {\it et al.}, Euro Phys. Lett. {\bf 34} 213
1265: (1996); R. Ambigapathy {\it et al.} Phys. Rev. Lett. {\bf 78} 3579
1266: q(1997); R. Kumar {\it et al.} Phys. Rev. Lett. {\bf 81}, 2578 (1998).
1267: 
1268: 
1269: \bibitem{baym61}G. Baym and L. P. Kadanoff, Phys. Rev. {\bf 124} 287
1270: (1961).
1271: 
1272: 
1273: \bibitem{flex} N.E. Bickers, D.J. Scalapino, and S.R. White,
1274: Phys. Rev. Lett. {\bf 62}, 961, (1989).
1275: 
1276: 
1277: \bibitem{kagan98} B. Kyung, E. G. Klepfish, and P. E. Kornilovitch,
1278: Phys. Rev. Lett. {\bf 80} 3109 (1998); M. Y. Kagan, R. Fr\'esard,
1279: M. Capezzali, and H. Beck, Phys. Rev. B {\bf 57}, 5995 (1998);
1280: 
1281: \bibitem{haussman} R. Haussman, Z. Phys. B {\bf 91}, 291 (1993).
1282: 
1283: 
1284: \bibitem{haughanamura} E. Hanamura, H. Haug, Physics Reports {\bf
1285: 33C}, 209, (1977).
1286: 
1287: \bibitem{tassone99xx} F. Tassone and Y. Yamamoto, Phys. Rev. B {\bf
1288: 59}, 10 830 (1999).
1289: 
1290: \bibitem{nambu60} Y. Nambu, Phys. Rev. B {\bf 117}, 648 (1960).
1291: 
1292: \bibitem{lindberg} M. Lindberg, and S.W. Koch, Phys. Rev. B {\bf 38},
1293: 3342 (1988).
1294: 
1295: \bibitem{piermarocchi99} C. Piermarocchi {\it et. al.} Solid State
1296: Commun. {\bf 112}, 433 (1999).
1297: 
1298: 
1299: \bibitem{kozlov96} J. Ding {\it et al.}, Phys. Rev. Lett. {\bf 69},
1300: 1707 (1992); V. Kozlov {\it et al.} Phys. Rev. B {\bf 54}, 13932
1301: (1996).
1302: 
1303: \bibitem{sirigu99} L. Sirgiu, D. Y. Oberli, L. Degiorgi, A. Rudra, and
1304: E. Kapon, Phys. Rev. B {\bf 61} R10575 (2000).
1305: 
1306: 
1307: \bibitem{dassarma00} S. Das Sarma and D. W. Wang,
1308: Phys. Rev. Lett. {\bf 84} 2010 (2000).
1309: 
1310: \bibitem{hannewald00} K. Hannewald, S. Glutch, and F. Bechstedt,
1311: Phys. Rev. B {\bf 62} 4519 (2000).
1312: 
1313: \bibitem{pieri98} P. Pieri and G. C. Strinati, eprint arXiv
1314: cond-mat/9811166 (1998) \end{references} %\end{multicols}
1315: \end{document}
1316: