cond-mat0303392/pra.tex
1: \documentstyle[aps,amsmath,latexsym,amssymb,epsfig,multicol,graphicx]{revtex}
2: \def\btt#1{{\tt$\backslash$#1}}
3: \begin{document}
4: 
5: \draft
6: \pagestyle{myheadings}
7: 
8: \newcommand{\comments}[1]{\hfill {\tt {eq:~#1}}} 
9: 
10: \def\breakon{\end{multicols}\widetext\vspace{-.2cm}
11: \noindent\rule{.48\linewidth}{.3mm}\rule{.3mm}{.3cm}\vspace{.0cm}}
12: 
13: \def\breakoff{\vspace{-.2cm}
14: \noindent
15: \rule{.52\linewidth}{.0mm}\rule[-.27cm]{.3mm}{.3cm}\rule{.48\linewidth}{.3mm}
16: \vspace{-.3cm}
17: \begin{multicols}{2}
18: \narrowtext}
19: 
20: \title{Polariton condensation and lasing in
21: optical microcavities - the decoherence driven crossover}  
22: 
23: \author{M. H. Szymanska, P. B. Littlewood, and  B. D. Simons}
24: \address{ Theory of Condensed Matter, Cavendish Laboratory, Cambridge
25: CB3 0HE, UK }
26: 
27: \date{\today}
28: 
29: \maketitle
30: 
31: \begin{abstract}
32: We explore the behaviour of a system which consists of a photon mode
33: dipole coupled to a medium of two-level oscillators in a microcavity
34: in the presence of decoherence. We consider two types of
35: decoherence processes which are analogous to magnetic and non-magnetic
36: impurities in superconductors. We study different phases of this
37: system as the decoherence strength and the excitation density is
38: changed. For a low decoherence we obtain a polariton condensate with
39: comparable excitonic and photonic parts at low densities and a
40: BCS-like state with bigger photon component due to the fermionic phase
41: space filling effect at high densities. In both cases there is a large
42: gap in the density of states. As the decoherence is increased the gap
43: is broadened and suppressed, resulting in a gapless condensate and
44: finally a suppression of the coherence in a low density regime and a
45: laser at high density limit. A crossover between these regimes is
46: studied in a self-consistent way analogous to the Abrikosov and Gor'kov
47: theory of gapless superconductivity \cite{abrikosov-gorkov}.
48: 
49: \end{abstract}
50: \pacs{42.50.Fx, 03.75.Gg, 42.50.Gy, 42.55.Ah}
51: 
52: 
53: \begin{multicols}{2}
54: \narrowtext
55: 
56: \section{Introduction}
57: 
58: The miniaturisation and improvement in the quality of optical cavities
59: in recent years led to the achievement of strong-coupling regime of
60: light-matter interaction in many physical systems.  The strong
61: coupling regime is characterised by well-developed coupled modes of
62: light and electronic excitations, called polaritons.  Polariton
63: splitting has been experimentally observed for atoms \cite{atompol},
64: quantum wells \cite{cavpol} and bulk excitons \cite{bulkcavpol},
65: excitons in organic semiconductors \cite{organicpol1,organicpol2},
66: exciton complexes \cite{chargedpol} and glass spheres. The
67: strong coupling regime has also been achieved for coupled Josephson
68: junctions \cite{JJ} in a microwave cavity.
69: 
70: With well-developed modes, that are sharp and have long lifetimes, a
71: natural question becomes the existence of coherent, condensed states.
72: A theoretically constructed polariton condensate is a mixture of
73: coherent state of light and coherent state of massive particles in the
74: media. It is characterised by two order parameters: the coherent
75: polarisation and the coherent photon field and exhibit a gap in the
76: excitation spectrum \cite{paul}. Since the polariton condensate would
77: be a source of coherent light the natural question arises how it is
78: different from and how it can be connected to the traditional laser.
79: 
80: The laser is a weak-coupling phenomenon: a coherent state of photons
81: created by stimulated emission from an inverted electronic population
82: due to strong pumping. The polarisation of the medium is heavily
83: damped and the atomic coherence is practically zero.  A coherent
84: photon field, oscillating at the bare cavity mode frequency, is the
85: only order parameter in the system \cite{laser}.
86: 
87: The crossover between a polariton condensate and a laser is sometimes
88: mistakenly attributed to an increase in density and a crossover from
89: bosonic (exciton) degrees of freedom to fermionic particle-hole
90: pairs. The absence of polariton splitting is associated with
91: disappearance of coherence, which is why experiments are concentrated
92: at low densities. In this work we show that the trend is an opposite
93: one - the condensate is more robust at high densities. One needs to
94: remember that polaritons, and so polariton splitting, are normal state
95: excitation and so disappearance of polariton splitting in the normal
96: state does not indicate what would happen if the system was condensed.
97: 
98: The issue of electronic coherence in polariton condensate is
99: independent of whether the excitations are more ``excitonic'' or more
100: like a two-component plasma (ionised electron-hole pairs), or indeed
101: whether or not there is saturation (phase-space filling) of the
102: electronic states.  It has been shown \cite{paul} that although the
103: increase in the density of electronic excitation leads to
104: nonlinearities in the polariton system which cause the collapse of the
105: splitting between the two polariton peaks in the normal state it does
106: not destroy condensation even at very high excitation densities.  The
107: saturation in the fermionic space forces the condensate to become more
108: photon like as the excitation density is increased and to have more
109: BCS-like character, but nevertheless the coherence in the media and
110: the gap in the excitation spectrum are present.  The change in the
111: density of the electronic excitations leads to the crossover between
112: the regime of BEC of polaritons and collective BCS-like state
113: 
114: This article will argue that the real enemy of condensation is
115: decoherence, not density, and that it is precisely decoherence which
116: drives the polariton condensate towards the laser regime. We show that
117: only the self-consistent inclusion of decoherence processes allows to
118: establish a crossover between an isolated condensate and a laser. The
119: widely used quantum Maxwell-Bloch (Langevin) equations with constant
120: decay rate for polarisation are not correct in a regime when the
121: coherent polarisation is large and the gap in the density of states is
122: present. We develop a self-consistent method analogous to the
123: Abrikosov and Gor'kov theory of gapless superconductors, which allows
124: us to study the stability of the polariton condensate at low
125: decoherence strength and to established the crossover to laser
126: behaviour as the decoherence is increased.
127: 
128: \section{Model}
129: \label{model}
130: 
131: The model we consider in this work is schematically shown in Figure
132: \ref{fig:model}.  It consists of a set of N two-level oscillators
133: dipole coupled to a single mode of electromagnetic field confined in a
134: cavity. This system is then subject to various decoherence, pumping
135: and damping processes.  These processes can be of a different physical
136: nature, depending on the material, but their exact details are not
137: that important for a general theory. They can be described, similarly
138: to laser theory, as baths of harmonic oscillators coupled to the
139: system in a way that gives the same effect as the real physical
140: environment.
141: 
142: \breakon
143: \begin{figure}[htbp]
144: 	\begin{center}
145: \includegraphics[width=16.2cm,angle=0]{model.eps} 
146: 	\end{center}
147: 	\caption{Sketch of the model studied in this work: the system
148: 	of two-level oscillators dipole coupled to a single cavity
149: 	mode interacting with various types of environment.}
150: \label{fig:model}
151: \end{figure}
152: \breakoff
153: 
154: 
155: Our model takes into account the major Coulomb interaction between the
156: electron and hole within the exciton, the phase-space filling effect,
157: disorder in the material (inhomogeneous broadening of excitonic
158: energies) and various types of decoherence effects.  However, it does
159: not include screening and Coulomb interactions between
160: excitons. Therefore, it gives a very good description of tightly
161: bound, Frenkel - type of excitons localised by disorder or bound on
162: impurities, molecular excitons in organic materials, atoms in the
163: solid state or Josephson junctions arrays in microwave
164: microcavity. And it gives only a qualitative description within a
165: mean-field approximation for other types of excitons like Wannier
166: excitons or excitons propagating in a sample (see Section
167: \ref{aplic}).
168: 
169: The model we use is the minimal required to describe the essential
170: physics. At this stage we do not intend to model any particular medium
171: with its complex interactions which would only make the general
172: picture less clear. More specific details of the particular medium
173: could, however, be included straightforwardly into the formalism. We 
174: will discuss this possibility in the Section~\ref{aplic}.
175: 
176: In the following, we consider a system comprised of an ensemble of $N$ 
177: two-level oscillators with an energy $\epsilon_j$ dipole coupled to a 
178: single cavity mode. The corresponding microscopic Hamiltonian takes the 
179: form
180: %
181: \begin{eqnarray}
182: &&\hat{H}_S = \omega_c\psi^\dagger\psi + \sum_{j=1}^N \epsilon_j
183: (b^\dagger_jb_j-a^\dagger_ja_j)\nonumber\\ 
184: &&\qquad\qquad + \sum_{j=1}^N \frac{g_j}{\sqrt{N}}(b^\dagger_ja_j\psi 
185: +\psi^\dagger a^\dagger_jb_j)
186: \label{HS}
187: \end{eqnarray}
188: %
189: where Fermionic operators $b_j$ and $a_j$ annihilate electrons in the upper 
190: and lower states respectively, while the Bosonic operator $\psi$ annihilates 
191: the photon. Here the sum extends over the possible sites $j$ where an 
192: exciton can be present (e.g. different molecules or localised states 
193: associated with a disorder potential). Matrix elements $g_j/\sqrt{N}$ 
194: describe the interaction of the photon with the two-level oscillators. 
195: 
196: Generally, the effect of the environment on the behaviour of the system 
197: can be modelled through the interaction of the internal degrees of freedom
198: with a bath. Taking into account different physical processes, the most
199: general coupling is of the form
200: %
201: \begin{eqnarray}
202: &&\hat{H}_{SB}=\sum_{k} g_{k}(\psi^\dagger d_{k}+d_{k}^\dagger \psi)
203: \nonumber\\
204: &&\qquad\qquad + \sum_{jk} \left[b^\dagger_j a_j 
205: (g^{\gamma_+}_{jk} c_{+,k}^\dagger+
206: g^{\gamma_-}_{jk} c_{-,k})+{\rm h.c}\right]\nonumber\\
207: &&\qquad\qquad + \sum_{jk} \Gamma^{(1)}_{jk}(b^\dagger_jb_j
208: +a^\dagger_ja_j)(c_{1,k}^\dagger +c_{1,k})\nonumber\\
209: &&\qquad\qquad + \sum_{jk} \Gamma^{(2)}_{jk}(b^\dagger_jb_j
210: -a^\dagger_ja_j)(c_{2,k}^\dagger +c_{2,k}),
211: \label{eq:HSB}
212: \end{eqnarray}
213: %
214: %
215: where 
216: %
217: \begin{eqnarray*} 
218: &&\hat{H}_{B}= \sum_{k} \left[\omega_{k} d_k^{\dagger}d_k +\omega_{+,k}
219: c^\dagger_{+,k} c_{+,k}+\omega_{-,k} c^\dagger_{-,k} c_{-,k}\right.\\
220: &&\qquad\qquad\qquad\qquad \left.+ \omega_{1,k} c^\dagger_{1,k} c_{1,k} 
221: +\omega_{2,k}c^\dagger_{2,k} c_{2,k}\right]
222: \end{eqnarray*}
223: %
224: describes the Hamiltonian of the bath. Here the different modes $k$ of the 
225: bath are indexed by (independent) Bosonic field operators $d_k$, $c_{+,k}$, 
226: $c_{-,k}$, $c_{1,k}$ and $c_{2,k}$. Here the first term in (\ref{eq:HSB}) 
227: describes the decay of the photon field from the cavity. Matrix elements
228: $g^{\gamma_+}_{jk}$ describe the incoherent pumping of two-level 
229: oscillators, while the matrix elements $g^{\gamma_-}_{jk}$ contain all
230: of the higher energy processes which destroy the electronic excitations 
231: such as the radiative decay
232: into photon modes different from the cavity mode. Apart from their dephasing
233: effect, together, these processes cause a flow of energy through the system. 
234: However, in the steady state, the total number of excitations
235: %
236: \begin{equation} 
237: \label{constrain}
238: \hat{n}_{ex}=\psi^\dagger\psi + \frac{1}{2}\sum_{j}(b^\dagger_jb_j-
239: a^\dagger_ja_j), 
240: \end{equation} 
241: %
242: the sum of photons and excited two-level oscillators, is constant. Finally, 
243: the third and fourth terms describe all those lower energy dephasing 
244: processes, such as
245: collisions and interactions with phonons and impurities, which conserve
246: the total number of excitations in the cavity. Such contributions can be 
247: divided into a part which act symmetrically ($\Gamma^{(1)}$) or 
248: antisymmetrically ($\Gamma^{(2)}$) on the upper and lower levels. Altogether 
249: these four terms contain all the essential mechanisms of decoherence.
250: 
251: To assimilate the effect of the different mechanisms of decoherence, we
252: will find it useful both intuitivelty as well as techincally (see later) 
253: to draw on an analogy between the Hamiltonian of the system and that of a
254: superconductor~\cite{keldysh1,keldysh2}. Referring to 
255: the states indexed by the Fermionic operators $b_j$ as `particle-like',
256: and those indexed by $a_j$ as `hole-like', $\hat{H}_S$ can be interpreted
257: as a BCS Hamiltonian for a superconductor with an imposed homogeneous 
258: `superconducting' order parameter $\psi$. With this analogy, it is clear
259: that the second and third terms of $\hat{H}_{SB}$~(\ref{eq:HSB}) affect 
260: a mechanism of `pair-breaking', while the fourth term is compatibile with
261: the symmetry of the Cooper pairs. The former act on the system as 
262: dynamically fluctuating magnetic impurities while the latter describe
263: dynamical fluctuations of a normal non-magnetic potential providing only 
264: an inhomogeneous broadening of the energies.
265: 
266: When the pumping and photon decay rates are high, the system would be 
267: driven out of equilibrium. However, if the thermalisation rate is in 
268: excess of the speed at which the system is pumped, an equilibrium 
269: assumption can be justified. In this work we will limit our considerations
270: to this regime focussing on the effect of decoherence on the equilibrium
271: system. Choosing both the pumping and decay rates to be small --- allowing
272: thermal equilibration --- their ratio can be used to fix the total 
273: excitation density $n_{ex}$. Nevertheless, within this quasi-equilibrium 
274: regime, both the third and fourth terms in (\ref{eq:HSB}) can, in fact, be 
275: arbitrary large since they do not couple to the total density $n_{ex}$. 
276: There is no restriction on the density of excitations nor on the decoherence 
277: rate.
278: 
279: Therefore, on this background, we will consider the total Hamiltonian
280: %
281: \begin{equation} 
282: \hat{H}=\hat{H}_S+\hat{H}_{SB}+\hat{H}_{B},
283: \label{H}
284: \end{equation} 
285: %
286: where
287: %
288: \begin{eqnarray}
289: &&\hat{H}_{SB} = \sum_{jk} \Gamma^{(1)}_{jk}(b^\dagger_jb_j
290: +a^\dagger_ja_j)(c_{1,k}^\dagger +c_{1,k})\nonumber\\
291: &&\qquad\qquad+\sum_{jk} \Gamma^{(2)}_{jk}(b^\dagger_jb_j
292: -a^\dagger_ja_j)(c_{2,k}^\dagger +c_{2,k}),
293: \label{Hmag}
294: \end{eqnarray}
295: %
296: includes only the includes only the decoherence mechanisms which conserve
297: $n_{ex}$, while 
298: %
299: %
300: \begin{eqnarray*} 
301: \hat{H}_{B}= \sum_{i=1,2} \sum_{k} \omega_{i,k} c^\dagger_{ik} c_{ik}.
302: \end{eqnarray*}
303: %
304: Although, in general, the coupling constants $\Gamma^{(i)}_{jk}$ can be site
305: dependent, for simplicity, we will suppose that the modes of the bath couple 
306: with equal strength to the two-level systems setting $\Gamma^{(i)}_{jk} 
307: \mapsto \Gamma^{(i)}_k$. Similarly, in the following, we will assume that the 
308: coupling of the two-level systems to the cavity photon is independent of 
309: the site index $j$, $g_j\mapsto g$.
310: 
311: When the interactions between the environment and the system are large, the
312: Hamiltonian above provides the basis of the standard theory of lasers (see,
313: e.g., Refs.~\cite{laser,haken}). At the same time, if we set the coupling
314: constants between the system and the environment to zero, the ground state
315: of the Hamiltonian $\hat{H}_S$ forms a polariton condensate. Thus by varying 
316: the magnitude of the coupling between the system and the baths the model
317: provides the means to move smoothly between an isolated condensate and
318: other phases driven by the decoherence. This provides a means to explore
319: the stability of the polariton condensate to interactions with the outside 
320: world at a small coupling strength, and to establish the connection between 
321: polariton condensation and lasers as the decoherence is increased.
322: 
323: The standard assumption (which in many cases is physically correct) is
324: that the environment leads to rapid dephasing of the exciton polarisation 
325: (i.e. $T_2$ is very short). It is thus generally assumed that the 
326: polarisation is very small, and that the coherent photon field is the
327: dominant order parameter. In such a case the baths can be averaged out 
328: before the exciton-photon interaction is studied. This in turn leads to the 
329: well-known quantum Maxwell-Bloch (Langevin) equations for the photon 
330: field and polarisation, essentially of the form
331: %
332: \begin{equation} 
333: \label{maxwell-bloch} \frac{d}{dt}
334: \langle a^\dagger b\rangle = i\langle[\hat{H}_{S}, a^\dagger b]\rangle
335: - \frac{1}{T_2} \langle a^\dagger b \rangle \;\;\;.  
336: \end{equation}
337: %
338: Crucially, Eq.~(\ref{maxwell-bloch}) can be derived by assuming that the
339: lifetime $T_2$ for polarisation of a \emph{single} two-level oscillator is
340: the same as that of the macroscopic ensemble of two-level oscillators (which
341: would correspond to introducing a separate bath for each two-level
342: system). However, in general Eq.~(\ref{maxwell-bloch}) is not correct
343: \cite{us0}. When $T_2$ is long, the macroscopic ensemble of two-level
344: oscillators can exist in a collective state characterised by a large
345: coherent polarisation, and the assumptions which lead to
346: (\ref{maxwell-bloch}) cannot be justified. Moreover, the constant
347: decay rate $1/T_2$ in (\ref{maxwell-bloch}) is a critical parameter;
348: even at arbitrarily small decoherence it leads to completely different
349: solutions from those in the absence of an environment (as shown in
350: Ref.~\cite{us0}). This criticality is however unphysical and arises only due
351: to approximations used in deriving the Langevin equations.
352: In fact, this conclusion seems not to be widely appreciated. Indeed, in a
353: relatively recent publication~\cite{wrong} similar decay constants in
354: the equations of motion for the coherently driven excitonic insulator
355: have been used, leading to the conclusion that the excitonic insulator
356: phase cannot exist for an arbitrary small decoherence.
357: 
358: One can gain some physical insight into this problem by considering the 
359: evolution of the density of states: The ideal condensate has a gap in 
360: the density of states which would still be present for small decoherence. 
361: It is evident that the coherent fields in this regime cannot be damped 
362: just by constant decay rates independent of frequency as there are no 
363: available states in which to decay. As the decoherence is increased this 
364: gap gets smaller and finally is completely suppressed causing the coherent 
365: fields to be strongly damped, as in lasers. In this regime 
366: Eq.~(\ref{maxwell-bloch}) is perfectly valid. However to be able to study 
367: a crossover from the fully phase coherent polariton condensate to a laser 
368: one needs to include the environment in a self-consistent way in which the 
369: possible gap in excitation spectrum and a large polarisation would be taken 
370: into account.
371: 
372: The analogy between superconductivity and the excitonic insulator was
373: first noticed and exploited in a pioneering work of Keldysh
374: \cite{keldysh1,keldysh2}, and explored later by others
375: \cite{zittart,paul}. The analogy between different types of
376: decoherence processes acting on polariton condensate, and the problem of
377: magnetic and potential impurities in superconductors suggests that similar
378: methods to that used by Abrikosov and Gor'kov (AG) in their theory of
379: gapless superconductivity~\cite{abrikosov-gorkov} can be useful in
380: studying the properties of polariton condensates in the presence of
381: decoherence. This analogy is however not exact and we will discuss a
382: few essential differences between our theory and the Abrikosov and Gor'kov
383: approach at the end of Section \ref{method}. An outline of the rest of
384: the paper is as follows: In section \ref{method} we provide details of
385: the method in the path integral formulation. In section \ref{gs} we
386: discuss ground state properties of the system in the presence of two
387: different types of dephasing while in section \ref{es} we study the
388: excitation spectrum and construct a phase diagram as a function of
389: excitation density, pair-breaking decoherence strength and the
390: inhomogeneous broadening of energy levels. In section \ref{cl} we
391: discuss a crossover between an isolated condensate and a laser and in
392: section \ref{gap} the magnitude of the energy gap. In section
393: \ref{aplic} we comment on the applicability of the model and the
394: method presented in this work and indicate a few directions in which
395: the model could be easily extended. Finally, in section \ref{exp} we
396: discuss recent experiments in the light of theoretical predictions
397: presented in this work and in section \ref{sum} we briefly summarise
398: the results.
399: 
400: \section{Path Integral Formulation}
401: \label{method}
402: 
403: To construct a theory of the coupled system, we will exploit an approach
404: based on the coherent state path integral. As well as providing 
405: access to the mean-field equations of the system, such a framework provides
406: the potential to explore the influence of fluctuation phenomena. Working
407: in the grand canonical ensemble, a chemical potential $\mu$ can be used
408: to fix the total number of excitations $n_{ex}$. The quantum partition 
409: function of the system ${\cal Z}={\rm Tr}\  e^{-\beta(\hat{H}-\mu
410: \hat{n}_{ex})}$ can be expressed as a coherent state path integral over 
411: Fermionic and Bosonic fields, 
412: %
413: \begin{equation} 
414: {\cal Z}=\int D[\bar{\psi}, \psi] \prod_j
415: D[\bar{\phi}_j,\phi_j] \prod_{k,i=1,2} D[\bar{c}_{i,k}, c_{i,k}] e^{-S}.
416: \label{partition}
417: \end{equation}
418: %
419: As with the Hamiltonian, it is convenient to separate the total action
420: as $S=S_S+S_{SB}+S_{B}$ where
421: %
422: \begin{eqnarray*}
423: S_S=S_\psi-\int_0^\beta d\tau \sum_j \bar\phi_j \hat{G}_{0j}^{-1} \phi_j
424: \end{eqnarray*}
425: %
426: with $S_\psi=\int_0^\beta d\tau \bar\psi(\partial_\tau+\omega_c-\mu)\psi$
427: denotes the action of the internal electron and photon degrees of freedom
428: of the system, while the action for the coupling of the system to the bath 
429: takes the form
430: %
431: \begin{eqnarray*}
432: &&S_{SB}=\int_0^\beta d\tau \sum_{i=1,2}\left[\sum_k \bar{c}_{ik}
433: (\partial_\tau +\omega_{ik}) c_{ik}\right.\\
434: &&\qquad\qquad\qquad\left.+\sum_k
435: \Gamma_k^{(i)}\rho^{(i)}(\bar{c}_{i,k}+ c_{i,k})\right].
436: \end{eqnarray*}
437: %
438: Combining the Fermionic fields of the two-level systems into a Nambu-like
439: spinor,
440: %
441: \begin{displaymath}
442: \bar{\phi}_j=\left( \begin{array}{c}
443: \bar{b}_{j}  \  \ \bar{a}_{j}
444: \end{array}
445: \right), 
446: \end{displaymath} 
447: %
448: the bare Green function assumes the matrix form
449: %
450: \begin{eqnarray*}
451: \hat{G}_{0j}^{-1}=-\partial_\tau\sigma_0-(\epsilon_j-\mu/2)\sigma_3-g\bar\psi
452: \sigma_- -g\psi\sigma_+\,.
453: \end{eqnarray*}
454: %
455: where $\sigma$ denote the Pauli spin matrices (with $\sigma_0\equiv\openone$)
456: which operate in the $(b,a)$ space (hereafter, loosely referred to as the 
457: `particle/hole' space). 
458: Finally, we have defined the symmetric and antisymmetric `densities' according 
459: to the relation $\rho(\tau)^{(1,2)}=\sum_j \bar{\phi}_j(\tau)\sigma_{(0,3)}
460: \phi_j(\tau)$.
461: 
462: Although a theory of the symmetric and antisymmetric processes can be 
463: developed in concert, for clarity we will present a detailed derivation
464: of the action of the `pair-breaking' decoherence processes imposed by 
465: $\Gamma^{(1)}$. Later, in Section \ref{saddle}, we will restore the 
466: decoherence processes affected by $\Gamma^{(2)}$. Thus, for now, we will 
467: use the following abbreviation $\Gamma_k^{(1)}\mapsto \Gamma_k$ dropping
468: the `channel' index. 
469: 
470: Being Gaussian in the Bosonic fields $c_{k}$, the degrees of freedom of the
471: bath can be integrated out leading to an effective interaction of the 
472: two-level systems which takes the form $\int D[\bar{c}_k,c_k] e^{-S_{SB}}
473: =e^{-S_{SB}'}$, where
474: %
475: \begin{equation}
476: \label{Sef}
477: S_{SB}'=\int_0^\beta d\tau d\tau' \rho(\tau) 
478: \sum_k \Gamma_k^2 D_k(\tau-\tau') \rho(\tau'),
479: \end{equation}
480: %
481: with $\hat{D}^{-1}_k=-\partial_\tau-\omega_k$ representing the free propagator 
482: of the environment. Transforming (\ref{Sef}) to the Fourier Matsubara 
483: frequency representation, and summing over the internal degrees of freedom 
484: of the bath,
485: %
486: \begin{equation}
487: \label{defgam}
488: -\sum_k \Gamma_k^2 D_k(i\nu_n)=f_{\Gamma}(i\nu_n),
489: \end{equation}
490: % 
491: where $D^{-1}_k(\nu_n)=i\nu_n-\omega_k$ and $\nu_n=2\pi n/\beta$, the 
492: induced interaction assumes the form
493: %
494: \begin{eqnarray*}
495: &&S_{SB}'=-\sum_{\nu_n} f_{\Gamma}(i\nu_n) \rho(i\nu_n)\rho(-i\nu_n)\\
496: &&=\int_0^\beta d\tau d\tau' f_{\Gamma}(\tau-\tau') \sum_{jj'} {\rm Tr}
497: \, \phi_j(\tau)\otimes\bar\phi_{j'}(\tau') \phi_{j'}(\tau') \otimes 
498: \bar\phi_j(\tau).
499: \end{eqnarray*}
500: %
501: In particular, it can be seen explicitly that the interaction with the 
502: environment introduces an effective quartic interaction between the
503: \emph{different} two-level systems. This contrasts with the Maxwell-Bloch
504: equations~(\ref{maxwell-bloch}) from which one can infer only a lifetime 
505: for excitations~\cite{us0}.
506: 
507: To develop a mean-field theory of the coupled system, it is helpful to
508: affect a Hubbard-Stratonovich decoupling of the interaction. Introducing 
509: the $2\times 2$ component matrix field $Q_{jj'}(\tau,\tau')$, which 
510: inherits the symmetry of the dyadic product $\phi_j(\tau)\otimes
511: \phi_{j'}(\tau')$, the interaction generated by the bath can be decoupled
512: as 
513: %
514: \begin{eqnarray*}
515: &&e^{-S_{SB}'}=\int DQ e^{-S_Q}\\
516: &&\times\exp\left[\int_0^\beta d\tau d\tau' f_\Gamma (\tau-\tau')
517: \sum_{jj'} \bar{\phi}_j(\tau) Q_{jj'}(\tau,\tau')\phi_{j'}(\tau')
518: \right],
519: \end{eqnarray*}
520: %
521: where
522: %
523: \begin{eqnarray*}
524: S_Q=\int_0^\beta d\tau d\tau' \sum_{jj'} f_\Gamma (\tau-\tau')
525: {\rm Tr}\ Q_{jj'}(\tau,\tau') Q_{j'j}(\tau',\tau)
526: \end{eqnarray*}
527: %
528: with trace taken in the particle-hole space. Combined with $S_S$, an
529: integration over the Fermionic degrees of freedom $\phi$ obtains the 
530: quantum partition function
531: %
532: \begin{equation} 
533: {\cal Z}=\int D[\bar{\psi}, \psi] \int DQ e^{-S}.
534: \end{equation}
535: %
536: where $S=S_Q+S_\psi-{\rm Tr}\ln \hat{G}^{-1}$, 
537: %
538: \begin{equation}
539: \label{Gtot}
540: G^{-1}_{jj'}(\tau,\tau')=G_{0j}^{-1}\delta_{jj'}\delta(\tau-\tau')-
541: f_\Gamma(\tau-\tau') Q_{jj'}(\tau,\tau').
542: \end{equation}
543: %
544: and the trace now runs over time and site indicies as well as the 
545: particle/hole space. At this level, the analysis is exact. 
546: 
547: \subsection{Saddle-point equations}
548: \label{saddle}
549: 
550: To develop the quantum partition function, further progress is possible 
551: only within a saddle-point approximation. Varying the action with respect
552: to $Q$, one finds that each electronic excitation is coupled to the average 
553: field created by all of the other excitations. In this sense, the 
554: saddle-point analysis corresponds to a mean-field treatment of the 
555: interaction between electronic excitations. This analysis becomes exact 
556: when there is a large number of electronic excitations coupled to a small 
557: number of field modes, since the fluctuations of the field are then 
558: negligible~\cite{paul}. The mean-field treatment of the interaction 
559: becomes exact in the thermodynamic limit when $N\to\infty$. The fluctuations 
560: above mean-field are of the order of $1/\sqrt{N}$. 
561: 
562: In the saddle-point approximation, it is assumed that the dominant
563: contribution to the quantum partition function (\ref{partition}) arises from
564: those configurations $\psi$ and $Q$ which minimise the total action.
565: Varying the action $S$ with respect to $Q$ one obtains the matrix equation
566: %
567: \begin{equation} 
568: Q_{jj'}(\tau,\tau')=\frac{1}{2} G_{jj'}(\tau,\tau'),
569: \label{saddlepointQ}
570: \end{equation}
571: %
572: while varying with respect to $\bar{\psi}$, the saddle-point equation 
573: takes the form
574: %
575: \begin{equation} 
576: (\partial_\tau+\omega_c-\mu)\psi(\tau)=\sum_j g {\rm Tr}\ G_{jj}(\tau,
577: \tau)\sigma_-.
578: \label{saddlepointF}
579: \end{equation}
580: %
581: Maintaining the analogy with the superconductor, the first equation 
582: which identifies $Q$ as the self-energy, describes the self-consistent 
583: Born approximation for the Green function, while the 
584: second equation represents the gap equation for the superconducting order
585: parameter. Substituting Eq.~(\ref{saddlepointQ}) into (\ref{Gtot}) the
586: saddle-point equation assumes the form of a Dyson equation
587: %
588: \begin{equation}
589: \label{Dyson}
590: G^{-1}_{jj'}(\tau,\tau')=G_{0j}^{-1}(\tau,\tau')\delta_{jj'}-
591: \Sigma_{jj'}(\tau,\tau'), 
592: \end{equation}
593: %
594: where 
595: %
596: \begin{eqnarray*}
597: \Sigma_{jj'}(\tau,\tau') =
598: \frac{1}{2}f_{\Gamma}(\tau-\tau')G_{jj'}(\tau,\tau')
599: \end{eqnarray*}
600: %
601: denotes the self-energy
602: 
603: 
604: Until now, we have focussed on the impact of the `pair-breaking' perturbation
605: affected by the matrix elements $\Gamma^{(1)}$. Consideration of the 
606: symmetric perturbation $\Gamma^{(2)}$ follows straightforwardly. In doing
607: so, it may be confirmed that the structure of the saddle-point equations
608: are maintained while the self-energy takes the form
609: %
610: \begin{eqnarray*}
611: \Sigma_{jj'}(\tau,\tau') =
612: \frac{1}{2}f_{\Gamma}(\tau-\tau')\sigma_3 G_{jj'}(\tau,\tau')\sigma_3
613: \end{eqnarray*}
614: 
615: In steady state one expects the solution of the saddle-point equations to
616: depend only on $\tau-\tau'$. In this case, a transformation to Matsubara
617: frequencies leads to the relation
618: %
619: \begin{equation}
620: \Sigma(i\nu_n) =
621: \frac{1}{2}\sum_{\nu_n'}f_{\Gamma}(i\nu_n')G(i\nu_n-i\nu_n').  
622: \label{selfen1}
623: \end{equation}
624: %
625: Generally, the solution of the saddle-point equation depends sensitively 
626: on the particular spectrum of decoherence $f_\Gamma$ and must be determined
627: self-consistently. However, an explicit solution to the saddle-point equation
628: can be established in various limits.
629: 
630: Generally, in order to determine $f_{\Gamma}$ from Eq.~(\ref{defgam}) one
631: can assume that coupling constants of the system to the bath $\Gamma_{k}$, 
632: as well as the bath density of states $N(\omega_{k})$, are continuous functions
633: of frequency. In this case, one can replace the summation over $k$ with an
634: integral over $\omega_{k}$ ($\Gamma_{k} \mapsto \Gamma(\omega_{k})
635: $, $\sum_{k} \mapsto \int d\omega_k N_k(\omega_k)$) whereupon
636: %
637: \begin{eqnarray*}
638: f_\Gamma(i\nu_n) = \int d\omega_k \frac{N_k(\omega_k)
639: \Gamma^2(\omega_k)}{-i \nu_n-\omega_k}.
640: \end{eqnarray*}
641: %
642: Now, in general $f_\Gamma$ will exhibit a particular frequency dependence 
643: determined by the density of states $N(\omega_k)$ of the bath and coupling
644: constant $\Gamma(\omega_k)$. However two special cases present themselves:
645: Firstly, if we assume that both $\Gamma(\omega_k)$ and $N(\omega_k)$ are 
646: largely independent of frequency over a wide range, one finds 
647: $f_\Gamma(i\nu_n) = i2\pi \Gamma^2 N$. This corresponds to a Markovian 
648: approximation to the bath in which $f_\Gamma(\tau)=i2\gamma^2\delta(\tau)$. 
649: In the second limit, if one assumes that the matrix elements and density
650: of states are concentrated at zero frequency, one has $f_\Gamma(i\nu_n)=
651: 2\gamma^2\delta_{\nu_n,0}$. Here, in the static limit, $f_\Gamma(\tau)=
652: 2\gamma^2$ is \emph{real}. Applied to the self-energy, the static limit 
653: leads to 
654: %
655: \begin{eqnarray*}
656: \Sigma_{jj'}(i\nu_n) = \gamma^2 G_{jj'}(i\nu_n), 
657: \end{eqnarray*}
658: %
659: while, in the Markovian approximation,
660: %
661: \begin{eqnarray*}
662: \Sigma_{jj'}(\tau,\tau) = i\gamma^2 G_{jj'}(\tau,\tau),
663: \end{eqnarray*}
664: %
665: (similarly for the perturbation $\Gamma^{(2)}$).
666: %
667: In these two limiting cases, the saddle-point equations admit a 
668: straightforward analytical solution. Here we present a detailed analysis
669: for the static limit.
670: 
671: Noting that photon field in the gap equation (\ref{saddlepointF}) couples 
672: only to the diagonal elements of the Green function, $G_{jj}$, this suggests
673: a mean-field Ansatz in which the matrix elements off diagonal in the $j$
674: space are taken to be zero, while the only time dependence of the
675: field is associated with oscillation at the chemical potential, $\mu$.
676: In this case, the coupled equations (\ref{saddlepointF}), (\ref{Dyson}) 
677: can now be solved.
678: 
679: To simplify the algebra we employ a similar mathematical trick to that
680: used by Abrikosov and Gor'kov in their theory of gapless superconductors. 
681: Since the overall phase of a coherent state is arbitrary, we can choose the 
682: mean-field $\langle \psi \rangle$ to be real and present the total Green
683: function $G$ (\ref{Dyson}) in the same form as the zero-order Green function 
684: $G_0$,
685: %
686: \begin{equation}
687: G^{-1}_{jj}(i\nu_n)=-i\tilde{\nu}_{j,n}-(\tilde{\epsilon}_j-\mu/2)\sigma_3-
688: g\langle \tilde{\psi}_j \rangle \sigma_1,
689: \label{G2}
690: \end{equation}
691: %
692: using the frequency dependent, renormalised $\tilde{\nu}_{j,n}$,
693: $\tilde{\epsilon}_j$ and $\langle \tilde{\psi}_j \rangle$.  Comparing
694: (\ref{Dyson}) with (\ref{G2}) we obtain for both type 1 and type 2
695: decoherence three equations determining the renormalised frequency,
696: energy and coherent photon field
697: %
698: \begin{align}
699: \label{eq:renorm-omega}
700: \tilde{\nu}_{j,n} &=\nu_n-\gamma_{1,2}^2 \frac{\tilde{\nu}_{j,n}}
701: {\tilde{\nu}_{j,n}^2+(\tilde{\epsilon}_j-\mu/2)^2 + g^2\langle 
702: \tilde{\psi}_j\rangle^2},\\
703: \label{eq:renorm-epsilon}
704: \tilde{\epsilon}_j &= \epsilon_j 
705: +\gamma_{1,2}^2 \frac{\tilde{\epsilon}_j}
706: {\tilde{\nu}_{j,n}^2+(\tilde{\epsilon}_j-\mu/2)^2 + g^2\langle 
707: \tilde{\psi}_j\rangle^2},\\
708: \langle \tilde{\psi}_j \rangle &=\langle \psi \rangle
709: \pm \gamma_{1,2}^2 \frac{g_j\langle \tilde{\psi}_j \rangle}
710: {\tilde{\nu}_{j,n}^2+(\tilde{\epsilon}_j-\mu/2)^2 + g^2\langle 
711: \tilde{\psi}_j\rangle^2}.
712: \label{eq:renorm-field}
713: \end{align}
714: %
715: while the gap equation (\ref{saddlepointF}) takes the form
716: %
717: \begin{equation} 
718: \langle\psi\rangle=\frac{g}{2(\omega_c-\mu)} \sum_{j} {\rm Tr} \ G_{jj}
719: \sigma_1.
720: \label{saddlepointF2}
721: \end{equation}
722: %
723: The average coherent polarisation of the medium can be determined from
724: the off-diagonal part of the Green's function $G$ (\ref{G2})
725: %
726: \begin{eqnarray}
727: &&\frac{1}{N}\langle \sum_j a^{\dagger}_jb_j \rangle = \langle P \rangle
728: \nonumber\\ 
729: &&\qquad = -\beta^{-1}\sum_{\nu_n,j} \frac{g\langle \tilde{\psi}_j \rangle}
730: {\tilde{\nu}_{j,n}^2+(\tilde{\epsilon}_j-\mu/2)^2 + g^2\langle 
731: \tilde{\psi}_j\rangle^2}.
732: \label{pol}
733: \end{eqnarray}
734: %
735: Thus, at the mean-field level, we can see from Eqs.~(\ref{saddlepointF2}) 
736: and (\ref{pol}) that the two order parameters, the coherent polarisation 
737: and the coherent photon field, are coupled according to the relation
738: %
739: \begin{equation}
740: \langle \psi \rangle= -\frac{g}{\omega_c-\mu} \langle P \rangle.
741: \label{eq:field_pol}
742: \end{equation}
743: %
744: The ratio between the two order parameters is determined
745: by the chemical potential, which in the steady state can be calculated
746: from Eq.~(\ref{constrain}). 
747: 
748: The number of electronic excitations, refered to later as inversion, can
749: be obtained from the diagonal elements of the Green's function
750: %
751: \begin{eqnarray}
752: &&\frac{1}{2}\langle \sum_j(b^{\dagger}_jb_j-a^{\dagger}_ja_j) \rangle
753: \nonumber\\ 
754: &&\qquad\qquad =-\beta^{-1}\sum_{\nu_n,j} \frac{2(\tilde{\epsilon}_j-\mu/2)}
755: {\tilde{\nu}_{j,n}^2+(\tilde{\epsilon}_j-\mu/2)^2 +
756: g^2\langle \tilde{\psi}_j \rangle^2}.
757: \label{inv}
758: \end{eqnarray}
759: 
760: Using Eqs. (\ref{eq:renorm-omega})-(\ref{eq:renorm-field}) we can
761: determine the renormalised parameters $\tilde{\nu}_{j,n}$, 
762: $\tilde{\epsilon}_j$ and $\langle \tilde{\psi}_j \rangle$ as a functions 
763: of the bare parameters $\nu_n$, $\epsilon_j$, $\langle \psi \rangle$ and 
764: $\gamma$. In the case of type 1 decoherence processes we obtain 
765: %
766: \breakon
767: \begin{equation}
768: \label{eq:last-field}
769: \langle \tilde{\psi}_j \rangle=\frac{\langle \psi \rangle}{2}+
770: \frac{\sqrt{2}\langle \psi \rangle}{4E_j}\sqrt{E_j^2-4\gamma_1^2
771: -\nu^2_n+\sqrt{-16\gamma^2_1E_j^2+(E_j^2+4\gamma^2_1+\nu^2_n)^2}},
772: \end{equation}
773: \breakoff
774: %
775: and 
776: %
777: \begin{align}
778: \label{eq:last-omega}
779: \tilde{\nu}_{j,n}&=\frac{\nu_n \langle \tilde{\psi}_j\rangle}
780: {2\langle \tilde{\psi}_j \rangle - \langle \psi \rangle}, \\
781: \label{eq:last-epsilon}
782: \tilde{\epsilon}_j&=\frac{\epsilon_j \langle \tilde{\psi}_j
783: \rangle}{\langle\psi \rangle},  
784: \end{align} 
785: %
786: while for the type 2 decoherence processes we have
787: %
788: \breakon
789: \begin{multline}
790: \label{eq:last-field2}
791: \langle \tilde{\psi}_j \rangle=\frac{\langle \psi \rangle}{2}+ \\
792: \frac{\sqrt{2}\langle \psi \rangle}{4(\nu_n^2+g^2\langle \psi\rangle^2)}
793: \sqrt{E_j^2-2(\epsilon-\mu/2)^2+4\gamma_1^2-\nu^2_n+
794: \sqrt{16\gamma^2_2(\nu_n^2+g^2\langle \psi
795: \rangle^2)+(E_j^2-4\gamma^2_1+\nu^2_n)^2}}, 
796: \end{multline}
797: \breakoff
798: %
799: and 
800: %
801: \begin{align}
802: \label{eq:last-omega2}
803: \tilde{\nu}_{j,n}&=\frac{\nu_n \langle \tilde{\psi}_j\rangle}{\langle\psi 
804: \rangle}, \\   
805: \label{eq:last-epsilon2}
806: \tilde{\epsilon}_j&=\frac{\epsilon_j \langle \tilde{\psi}_j\rangle}
807: {2\langle \tilde{\psi}_j \rangle - \langle \psi \rangle},
808: \end{align}
809: %
810: where for both cases
811: %
812: \begin{equation}
813: \label{eq:Equasip}
814: E_j=\sqrt{(\epsilon_j-\mu/2)^2+g^2\langle \psi \rangle^2}.
815: \end{equation}
816: %
817: Substituting Eqs.~(\ref{eq:last-field})-(\ref{eq:last-epsilon}) or
818: (\ref{eq:last-field2})-(\ref{eq:last-epsilon2}) into Eq.~(\ref{pol}),
819: summing over the Matsubara frequencies, and using Eq.~(\ref{eq:field_pol}) 
820: we can determine the ground state coherent polarisation $\langle P \rangle$ 
821: and the coherent photon field $\langle \psi \rangle$ as functions of the 
822: system parameters $\epsilon$, $\omega_c$, decoherence parameter 
823: $\gamma_{1,2}$ and chemical potential $\mu$. The chemical potential can be 
824: then obtained from Eq.~(\ref{constrain}). The integrals over the Matsubara 
825: frequencies at zero temperature in (\ref{pol}) and (\ref{inv}) as well as the
826: determination of the chemical potential from (\ref{constrain}) have to
827: be performed numerically.
828: 
829: \subsection{Density of states}
830: 
831: The excitation spectrum, i.e the density of states, can be obtained
832: from the diagonal part of the Green function on the real frequency
833: axis. Considering the analytical continuation of $G_{jj}(i\nu_n) \to
834: \bar G_{jj}(\nu)$, where $\nu_n$ and $\nu$ are the Matsubara and the
835: real frequencies respectively, and thus using the usual substitution
836: $i\nu_n \to -\nu +i \delta$, we obtain the relationship between the
837: Green function $G_{jj}(i \nu_n)$ and the density of states $A(\nu)$
838: %
839: \begin{equation}
840: A(\nu)=\sum_j\lim_{\delta \to 0^+}{\rm Im} G_{jj}(-\nu+i\delta+\mu),
841: \end{equation}
842: %
843: which has the following form
844: %
845: \begin{equation}
846: \label{eq:spectrum}
847: A(\nu)=\sum_j {\rm Im} \frac{\tilde{\nu}_j+(\tilde{\epsilon}_j-\mu/2)}
848: {\tilde{\nu}_j^2+(\tilde{\epsilon}_j-\mu/2)^2 +
849: g^2\langle \tilde{\psi}_j \rangle^2}.
850: \end{equation}
851: %
852: Generally, $\tilde{\nu}_j$, $\tilde{\epsilon}_j$ and $\langle
853: \tilde{\psi}_j \rangle$ are functions of $\nu$, $\epsilon_j$, $\langle
854: \psi \rangle$ and $\gamma$ which can be determined from Eqs.
855: (\ref{eq:last-field})-(\ref{eq:last-epsilon}) or
856: (\ref{eq:last-field2})-(\ref{eq:last-epsilon2}) by the following
857: substitution $i\nu_n \to -\nu +i \delta$. It can be shown 
858: from Eq.~(\ref{eq:spectrum}) that the system of 
859: two-level oscillators with uniform energies, $\epsilon_j=\epsilon$, in
860: the presence of the type 1 processes has a gap, $\Delta$, in the
861: density of states of magnitude
862: %
863: \begin{equation}
864: \Delta=2\sqrt{(\epsilon-\mu/2)^2+g^2\langle \psi \rangle^2}-4\gamma_1.
865: \label{eq:gap}
866: \end{equation}
867: %
868: At very high excitation densities the gap is proportional to the
869: coherent field amplitude $\Delta \approx 2g\langle \psi \rangle
870: -4\gamma_1$. At very low excitation densities, when $\langle \psi
871: \rangle \to 0$, we recover conventional polaritons for
872: which the chemical potential is $2\epsilon-\mu \to g$ and the gap
873: $\Delta \to g-4\gamma_1$.
874: 
875: 
876: The major difference between this work and the AG theory
877: \cite{abrikosov-gorkov} is that the system studied here has two order
878: parameters connected through the chemical potential which needs to be
879: determined. We use a different form of the density of states for the
880: two-level oscillators than in their theory. Instead of a flat
881: distribution of energies from $-\infty$ to $+\infty$ used in the AG
882: method we first perform the calculations for the degenerate case where
883: all two-level oscillators have the same energy $\epsilon$ and then we
884: use a realistic Gaussian distribution of energies, present in the real
885: microcavities. To account for these differences we need to include the
886: additional, third equation for renormalised $\tilde{\epsilon}$ not
887: present in the original Abrikosov and Gor'kov method and the
888: constraint equation for $n_{ex}$. In the Abrikosov and Gor'kov theory
889: they consider free propagating electrons with momentum {\bf k} over
890: which all the summations are performed. In our model of the localised
891: two-level systems the summations are performed over the sites where
892: the two-level oscillators can be present. Dynamic impurities can
893: easily be included in our formalism.
894: 
895: 
896: To perform the calculations we rescale the coherent fields by
897: $\sqrt{N}$ and consequently the inversion and the number of
898: excitations by $N$ introducing the excitation density
899: $\rho_{ex}=n_{ex}/N$. In this terminology the minimum $\rho_{ex}=-0.5$
900: corresponds to no photons and no electronic excitations in the
901: system. The condition $\rho_{ex}=0.5$ in the absence of photons would
902: correspond to all two-level oscillators in excited states, thus to the
903: maximum inversion.
904: 
905: We calculate first the ground state coherent field $\langle \psi
906: \rangle$, the coherent polarisation $\langle P \rangle$, the inversion
907: and the chemical potential as functions of the decoherence strength
908: $\gamma$ and the excitation density $\rho_{ex}$ for different
909: distributions of excitonic energies. Then we study the excitation
910: spectrum of the system for different regimes. The ground state
911: properties and the excitation spectrum allow us to obtain a phase
912: diagram for different excitation densities and decoherence
913: strengths. We consider the influence of both type 1 and type 2
914: decoherence processes as well as inhomogeneous broadening of exciton
915: energies.
916: 
917: \section{The Ground State --- Coherent Fields}
918: \label{gs}
919: 
920: \subsection{Type 1 (Pair-Breaking) Decoherence Processes}
921: \label{type1}
922: 
923: To examine the ground state properties of the system in the presence
924: of the type 1 decoherence processes we study the mean value of the
925: annihilation operator of the field and the polarisation. This mean is
926: non-zero only in a coherent state.  Figure \ref{fig:phot} (upper
927: panel) shows the behaviour of the coherent part of the photon field
928: $\langle \psi \rangle$ as the decoherence strength $\gamma$ is changed
929: for different excitation densities $\rho_{ex}$ and different
930: inhomogeneous broadenings of exciton energies.
931: 
932: For small values of $\gamma/g$, up to some critical value
933: $\gamma_{C1}$, $\langle \psi\rangle$ is practically unchanged while
934: for $\gamma/g>\gamma_{C1}$ the coherent field is damped quite rapidly
935: with increasing decoherence. This critical value of the
936: decoherence strength, $\gamma_{C1}$ is proportional to $\rho_{ex}$,
937: suggesting that for higher excitation densities the system is more
938: resistant to dephasing. At low excitation densities, where
939: $\rho_{ex}<0$, there is a second critical value of the decoherence
940: strength, $\gamma_{C2}$, where both coherent fields are sharply damped
941: to zero. As the excitation density is increased, precisely at
942: $\rho_{ex}=0$, $\gamma_{C2}$ diverges and does not exist
943: for $\rho_{ex}>0$ --- coherent fields although reduced are never
944: completely suppressed.
945: 
946: The behaviour of the electronic inversion, which is a measure of the
947: excitonic density given by Eq.~(\ref{inv}), is presented in
948: Fig. \ref{fig:phot} (middle panel). In our terminology an inversion
949: of -0.5 corresponds to no excitons while an inversion equal to zero means 
950: that the excitonic system is half occupied (roughly 0.5 per Bohr radius in a
951: model where double occupation of excitonic sites is not
952: allowed). Inversion of 0.5 would then correspond to 1 exciton per Bohr
953: radius. In this region of the decoherence strength where $\langle \psi
954: \rangle$ is damped, the inversion increases. At low excitation
955: densities ($\rho_{ex}<0$) the inversion approaches $\rho_{ex}$ for
956: $\gamma/g=\gamma_{C2}$ and stays constant as $\gamma$ is further
957: increased. At high excitation densities ($\rho_{ex}>0$) the inversion
958: asymptotically approaches zero with increasing dephasing.  The maximum
959: value of electronic inversion, for any exciton density and decoherence
960: strength, is zero, which corresponds to a half filled excitonic
961: system. This is a consequence of our assumption of thermal equilibrium in the 
962: exciton-photon system.
963: 
964: The ratio of coherent polarisation to coherent field $\langle
965: P \rangle/\langle \psi \rangle$ is presented in Fig. \ref{fig:phot}
966: (lower panel). For an isolated system, where $\gamma=0$, this ratio
967: depends on the excitation density. The condensate becomes more
968: photon like as $\rho_{ex}$ is increased due to the phase space filling
969: effect. For finite $\gamma$, at a given excitation density, this ratio
970: decreases with increasing $\gamma$ meaning that the coherent
971: polarisation is more heavily damped than the coherent photon field by
972: the type 1 decoherence processes. At $\rho_{ex}<0$ this ratio becomes
973: undefined for $\gamma/g>\gamma_{C2}$ when both coherent fields vanish.
974: 
975: In order to study the system of realistic, inhomogeneously broadened
976: two-level oscillators we have replaced the summations over sites with
977: integrals over the energy distribution. We have assumed this
978: distribution to be a Gaussian with mean $\epsilon_0$ and variance
979: $\sigma $. Our results, presented as dashed and dashed-dotted lines in
980: Fig \ref{fig:phot} show that a Gaussian broadening of energies does
981: not make any qualitative difference to the degenerate case. The coherent
982: fields and the critical values of decoherence strength, $\gamma_{C1}$
983: and $\gamma_{C2}$ are, as expected, slightly smaller than in the
984: degenerate case but all the regimes are analogous.
985: 
986: \begin{figure}[htbp]
987: 	\begin{center}
988: \includegraphics[width=8.8cm,angle=0]{ph_inv_ratio.eps}
989: 	\end{center} 
990: 	\caption{Coherent photon field $\langle \psi \rangle$ (upper
991: 	panel), inversion (middle panel) and ratio between coherent
992: 	photon field and coherent polarisation $\langle P
993: 	\rangle/\langle \psi \rangle$ (lower panel) as functions of
994: 	the pair-breaking decoherence strength, $\gamma/g$ for
995: 	different excitation densities, $\rho_{ex}$ and variances of
996: 	inhomogeneous broadening $\sigma=0$ (solid lines), $\sigma=0.5$
997: 	(dashed lines) and $\sigma=1.0$ (dotted lines). }
998: 
999: \label{fig:phot}
1000: \end{figure}
1001: %
1002: 
1003: \subsection{Type 2 (Non-Pair-Breaking) Decoherence Processes}
1004: \label{type2}
1005: 
1006: \begin{figure}[htbp]
1007: 	\begin{center} 
1008: \includegraphics[width=8.8cm,angle=0]{magnonmag2.eps}
1009: 	\end{center} 
1010: 	\caption{Comparison between the influence of a
1011: 	pair-breaking (solid line) and a non-pair-breaking (dotted
1012: 	line) decoherence processes on $\langle \psi \rangle$ (upper
1013: 	panel), inversion (middle panel) and $\langle P
1014: 	\rangle/\langle \psi \rangle$ (lower panel) for three different
1015: 	values of $\rho_{ex}$.}
1016: \label{fig:photnonm}
1017: \end{figure}
1018: 
1019: We now repeat the analysis for type 2 decoherence processes
1020: (\ref{Hmag}), which act in an exactly the same way on the upper and
1021: the lower levels of the two-level oscillator. Recall that such processes
1022: mirror the effects of non-magnetic impurities in the superconductor.
1023: 
1024: In Fig. \ref{fig:photnonm} we present for comparison the coherent
1025: photon field (upper panel), the inversion (middle panel), and the
1026: ratio between the coherent polarisation and the coherent photon field
1027: (lower panel) in the presence of the type 1 (solid line) and the type
1028: 2 (dotted line) decoherence processes for three excitation densities
1029: $\rho_{ex}=-0.4$, $\rho_{ex}=-0.2$ and $\rho_{ex}=0.2$. It is evident
1030: that the singular behaviour at $\gamma_{C1}$ and $\gamma_{C2}$
1031: discussed for the type 1 decoherence processes is not present for the
1032: case of the type 2 processes. With an increase of the decoherence
1033: strength from zero the coherent fields are slightly reduced and they
1034: slowly decrease, asymptotically approaching zero at low excitation
1035: densities ($\rho_{ex}<0$) or a constant value at high densities
1036: ($\rho_{ex}>0$). Although both coherent fields are reduced the
1037: behaviour of their ratio strongly depends on the excitation
1038: density. We will show in Section \ref{nonmagnetic} that the type 2
1039: processes give rise to the broadening of energies and then the
1040: behaviour of the ratio $\langle P \rangle/\langle \psi \rangle$
1041: depends on the position of the chemical potential with respect to the
1042: energy distribution.  The type 2 processes can make the condensate
1043: more photon or more exciton like depending on the parameters of the
1044: system. The two different cases are presented in
1045: Fig. \ref{fig:photnonm} (lower panel).
1046: 
1047: \section{Excitation Spectrum and the Phase Diagram}
1048: \label{es}
1049: To understand the behaviour of the coherent fields we study the
1050: excitation spectrum of the system in different regimes.  The density
1051: of states of the system with uniform energy distribution for six
1052: different values of $\gamma$ at low excitation density
1053: $\rho_{ex}=-0.4$ is presented in Fig. \ref{fig:green-0.4} while at
1054: high excitation density $\rho_{ex}=-0.2$ in Fig. \ref{fig:green0.2}.
1055: %
1056: 
1057: 
1058: \subsection{Type 1 (Pair-Breaking) Decoherence Processes}
1059: In the absence of decoherence (Fig. \ref{fig:green-0.4} a and
1060: \ref{fig:green0.2} a) we have two sharp peaks at two quasi-particle
1061: energies, $\pm E$, given by equation (\ref{eq:Equasip}) for the
1062: degenerate case ($E_j=E$). As $\gamma$ increases these two peaks broaden,
1063: which causes a decrease in the magnitude of the energy gap
1064: (Fig. \ref{fig:green-0.4} b, c and \ref{fig:green0.2} b, c, d ). The
1065: magnitude of the energy gap in the degenerate case is equal to
1066: $2E-4\gamma$, which is given in more detail in equation
1067: (\ref{eq:gap}). Finally, precisely at $\gamma_{C1}$ (shown in Fig.
1068: \ref{fig:phot}), these two broadened peaks join together and the gap
1069: closes (Fig. \ref{fig:green-0.4} d and \ref{fig:green0.2} e). When the
1070: decoherence strength is increased further (Fig. \ref{fig:green-0.4} e)
1071: these two peaks overlap more and the shape of the gapless density of
1072: states changes. For $\gamma/g>\gamma_{C2}$ at low densities the
1073: coherent fields are suppressed, thus Fig \ref{fig:green-0.4} f shows
1074: the normal state density of states in the absence of coherence.  Finally, 
1075: at high densities (Fig. \ref{fig:green0.2} f), as one would expect for a 
1076: laser system, the coherent field is present without a sign of a gap.
1077: 
1078: %
1079: \begin{figure}[htbp]
1080: 	\begin{center}
1081: \includegraphics[width=8.8cm,angle=0]{green-0.4.eps} 
1082: 	\end{center}
1083: 	\caption{Density of states for $\rho_{ex}=-0.4$ and different
1084:   	decoherence strengths, $\gamma/g$ for a pair-breaking (solid
1085:   	line) and a non-pair-breaking (dotted line) decoherence
1086:   	processes.}
1087: \label{fig:green-0.4}
1088: \end{figure}
1089: 
1090: Fig \ref{fig:greenb} shows the density of states for the system of the
1091: inhomogeneously broadened two-level oscillators with standard
1092: deviation $\sigma=0.5g$ for different values of the type 1 decoherence
1093: strength at two different excitation densities.  The broadening of the
1094: density of states and the suppression of the energy gap can be
1095: observed as $\gamma$ is increased. The last curve in the Fig
1096: \ref{fig:greenb} (left panel) shows the normal state density 
1097: where the coherent fields are suppressed. In the Fig \ref{fig:greenb}
1098: (right panel) where the coherent fields are present for all the values
1099: of $\gamma$ the last curve has no sign of a gap.
1100: 
1101: There are clearly three different phases depending on the decoherence
1102: strength, $\gamma$ (homogeneous broadening), inhomogeneous broadening,
1103: $\sigma$ and the excitation density $\rho_{ex}$. In the Fig.
1104: \ref{fig:phase} we present a phase diagram for the system.  The phase
1105: boundaries are defined by $\gamma_{C1}$ and $\gamma_{C2}$ for
1106: different values of $\rho_{ex}$ and $\sigma$.
1107: 
1108: Below the white surface, for small decoherence, we have a phase in
1109: which both coherent fields and an energy gap in the density of states
1110: are present. In this region coherent fields are protected by the
1111: energy gap and remain practically unchanged while the energy gap
1112: narrows as the decoherence is increased. At low densities within this
1113: phase we have a BEC of polaritons with the electronic and photonic
1114: parts comparable in size. At high densities we have a BCS-type of
1115: condensate with photon component increasing with the excitation
1116: density. Despite the predominantly photon-like character of this phase
1117: at very high densities, the coherence of the medium is still large and
1118: this phase can be distinguished from a laser by the presence of a gap in
1119: the density of states (Fig \ref{fig:greenb} - right panel). The
1120: crossover between high and low densities is a smooth evolution and
1121: there are no rapid phase transitions.
1122: 
1123: %
1124: \begin{figure}[htbp]
1125: 	\begin{center}
1126: \includegraphics[width=8.8cm,angle=0]{green0.2.eps}
1127: 	\end{center}
1128: 	\caption{Density of states for $\rho_{ex}=0.2$ and different
1129:   	decoherence strengths, $\gamma/g$ for a pair-breaking (solid
1130:   	line) and a non-pair-breaking (dotted line) decoherence
1131:   	processes.}
1132: \label{fig:green0.2}
1133: \end{figure}
1134: 
1135: 
1136: 
1137: Inhomogeneous broadening has some influence on this phase mainly at low
1138: excitation densities. The critical decoherence $\gamma_{C1}$, which
1139: defines a boundary between gapped and gapless phases, decreases with an
1140: increase in the inhomogeneous broadening $\sigma$. At high excitation
1141: densities inhomogeneous broadening has very weak influence on the
1142: system.   
1143: %
1144: 
1145: Between the white and gray surfaces there is a phase where the
1146: coherent fields are present without an energy gap in the density of
1147: states. This phase exists for all values of the decoherence parameter
1148: at high excitation densities $\rho_{ex} >0$. The coherent fields are no
1149: longer protected by the gap and get reduced as the decoherence is
1150: increased. The coherent polarisation is more heavily damped than the
1151: coherent field.  Within this phase we have at low densities a gapless,
1152: light-matter condensate, analogous to gapless superconductivity, whilst
1153: at very high densities we have essentially the laser system.
1154: Finally, above the dashed line there is a phase where the coherent
1155: fields are completely suppressed.
1156: 
1157: 
1158: %
1159: \breakon
1160: \begin{figure}[htbp]
1161: 	\begin{center}
1162: \includegraphics[width=16.2cm,angle=0]{green3D.eps}
1163: 	\end{center}
1164: 	\caption{Density of states for the Gaussian broaden case
1165: 	with $\sigma=0.5$ for different values of the pair-breaking
1166: 	decoherence strength at $\rho_{ex}=-0.3$ (left panel) and
1167: 	$\rho_{ex}=0.2$ (right panel). }
1168: \label{fig:greenb}
1169: \end{figure}    
1170: \breakoff
1171: %
1172: 
1173: %
1174: \begin{figure}[htbp]
1175: 	\begin{center}
1176: \includegraphics[width=8.8cm,angle=0]{phase.eps}
1177: 	\end{center}
1178: 	\caption{Phase diagram. The phase boundaries $\gamma_{C1}$ and
1179: 	$\gamma_{C2}$ as marked in Fig \ref{fig:phot}.  }
1180: \label{fig:phase}
1181: \end{figure}
1182: %
1183: 
1184: \subsection{Type 2 (Non-Pair-Breaking) Decoherence Processes}
1185: \label{nonmagnetic} 
1186: 
1187: The density of states in the presence of type 2 processes is shown in
1188: Figs \ref{fig:green-0.4} and \ref{fig:green0.2} (dotted lines). It can
1189: be seen that, although the two quasiparticle peaks get very broad,
1190: the gap is only slightly affected by the type 2 processes and is
1191: always present. Even for much larger values of $\gamma$ than presented
1192: in Figs \ref{fig:green-0.4} and \ref{fig:green0.2} the gap is not
1193: suppressed. The gap in the density of states is present until the
1194: coherent fields get completely suppressed. At high excitation
1195: densities ($\rho_{ex}>0$) coherent fields are always present and thus
1196: the density of states will have a gap for all values of the
1197: decoherence strength.
1198: 
1199: The type 2 processes give a similar effect as the inhomogeneous
1200: broadening of energy levels in the case of an isolated system (see Ref.
1201: \cite{paul}). The difference is that the density of states in the
1202: presence of the type 2 processes has sharp boundaries. This is a
1203: result of the method which is used to perform the calculations. It can
1204: be seen that if the bath operators for the type 2 processes,
1205: $c_2$'s in expression (\ref{Hmag}), were just numbers, this
1206: term could have been included into the second term in equation
1207: (\ref{HS}) and would had just given a random $\epsilon_i$. In our
1208: calculations the $c_2$'s are operators and we use a
1209: self-consistent Born approximation, which is not exact, and does
1210: not correctly reproduce the tail of the distribution. Thus, because 
1211: of the way the
1212: calculations are done, the density of states produced always has sharp
1213: boundaries.
1214: 
1215: In the Abrikosov and Gor'kov theory, due to the flat density of states
1216: used in the calculations, the non-magnetic impurities do not influence
1217: the superconducting state at all. In the case of degenerate or realistic,
1218: Gaussian distributed energies of the two level oscillators the type 2
1219: processes have some quantitative influence on the coherent fields and
1220: the gap but cannot cause any phase transitions.
1221: 
1222: \section{Crossover to Laser}
1223: \label{cl}
1224: 
1225: The features of the second phase (between the white and the gray
1226: surfaces) in Fig. \ref{fig:phase} at high $\rho_{ex}$ are essentially
1227: the same as those of the laser system. The laser operates in the
1228: regime of a very strong decoherence, comparable with the light-matter
1229: interaction itself. In the presence of such a large decoherence laser
1230: action can be observed only for a sufficiently large excitation
1231: density.  The coherent polarisation in a laser system is much more
1232: heavily damped than the photon field and the gap in the density of
1233: states is not observed. Thus the laser is a regime of our system for
1234: very large $\rho_{ex}$ and $\gamma$.
1235: 
1236: Laser theories, due to the approximations on which they are based, can
1237: only be valid in a regime where the gap in the density of states is
1238: suppressed and thus for a large decoherence. At the time when these
1239: theories were proposed they were deemed sufficient as most lasers
1240: operate in such a regime. Miniaturisation and improvements in the
1241: quality of optical cavities in recent years can lead to a large
1242: suppression of decoherence in laser media. For small decoherence and
1243: very small pumping in comparison to decay processes, when $\rho_{ex}$
1244: is much smaller then $0$, the laser theories would predict a lack of
1245: coherence while the real ground state of the system would be a more
1246: matter-like condensate. Thus an extension of laser theories to account
1247: for the gap in the excitation spectrum and coherence in a media is
1248: necessary. In our theory the laser emerges from the polariton
1249: condensate at high densities when the gap in the density of states
1250: closes for large decoherence and thus is analogous to a gapless
1251: superconductor.
1252: 
1253: When both the gap and the coherence in a media are taken into account,
1254: in contrast to a traditional laser, the coherent photon field can be
1255: present without the population inversion in the media. The polariton
1256: condensate is thus an example of a laser without inversion.
1257: 
1258: However, it has to be pointed out that there is no formal distinction
1259: between a laser and a Bose condensate of polaritons.  In the laser,
1260: coherence in the medium (manifested by the coherent polarisation)
1261: although small, is not completely suppressed so the laser can be seen
1262: as a gapless condensate with a more photon-like character. One of the
1263: possible distinctions between BEC and laser could be an existence of
1264: an energy gap in the excitation spectrum.
1265: 
1266: \section{The gap}
1267: \label{gap}
1268: \begin{figure}[htbp]
1269: 	\begin{center}
1270:  \includegraphics[width=8.8cm,angle=0]{del00.eps}	
1271: 	\end{center}
1272: 	\caption{Energy gap $\Delta_0$ in the units of g in the
1273: 	absence of decoherence as a function of excitation density
1274: 	$\rho_{ex}$, and Gaussian broadening, $\sigma$.}
1275: \label{fig:del0}
1276: \end{figure}
1277: 
1278: In the degenerate case the magnitude of the gap in the density of states
1279: is given by (\ref{eq:gap})  If we consider an inhomogeneous broadening
1280: of excitonic energies with a Gaussian distribution then there will always
1281: be some oscillators with energies which make the term $(\epsilon -
1282: \mu/2)^2$ vanish and thus, strictly, the magnitude of the gap would be
1283: $2g\langle \psi \rangle-4\gamma_1$ at all excitation densities. This
1284: means that the gap would vanish at $\langle \psi \rangle \to 0$ in
1285: contrast to degenerate case in which its magnitude would be $g - 4
1286: \gamma_1$. However, the contribution of these states, arising only from
1287: the tails of the Gaussian distribution, would be very small
1288: essentially leading to a soft gap of the magnitude of $g - 4
1289: \gamma_1$ at very low densities as in degenerate case. 
1290: 
1291: We call $\Delta_0$ the gap in the absence of decoherence when
1292: $\gamma_1=0$. Omitting the case when $n_{exc}=0$, discussed above, we
1293: calculate numerically the magnitude of the gap, in the units of the
1294: dipole coupling $g$, for a wide range of excitation densities and
1295: inhomogeneous broadening $\sigma>0$. $\Delta_0$ very much depends on
1296: the excitation density. The Gaussian broadening of energies decreases
1297: the gap at low densities but has almost no influence at high
1298: excitation densities. In the presence of the decoherence (homogeneous
1299: broadening) the actual gap $\Delta$ would be
1300: %
1301: \begin{equation} 
1302: \Delta = \Delta_0-4\gamma_1
1303: \label{totgap}
1304: \end{equation} 
1305: %
1306: Fig. \ref{fig:del0} and equation (\ref{totgap}) can be used to
1307: estimate the magnitude of the gap in particular experimental systems
1308: in which the values of broadenings, density and the dipole coupling is
1309: known.  Some estimates of the magnitude of the energy gap in
1310: particular materials and conditions were reported in a brief report
1311: \cite{us}.
1312: 
1313: \section{Applicability of the model and the method}
1314: \label{aplic}
1315: 
1316: As already stated in Section \ref{model}, the model of two-level
1317: systems interacting only via cavity photon field gives a reliable
1318: description of  tightly bound, Frenkel-type excitons localised
1319: by disorder or bound on impurities, molecular excitons in organic
1320: materials, atoms in the solid state or Josephson junctions arrays in
1321: microwave microcavities. In the classical limit $lim_{N \rightarrow
1322: \infty} (\langle n_{ex}\rangle /N) = \rho \rightarrow const.$ it has 
1323: an exact solution of the mean-field form
1324: %
1325: \begin{equation}
1326: \label{varwfn3}
1327: |\lambda,u,v\rangle = e^{\lambda \psi^\dagger }
1328: \prod_{j} ( v_{j} b_j^{\dagger} + u_{j} a_j^{\dagger} )|0\rangle.
1329: \end{equation}
1330: %
1331: However, the case of Wannier excitons propagating in the sample,
1332: present in some clean semiconductors (for example GaAs quantum wells), is
1333: conceptually not very different. The sums over sites in the microscopic
1334: Hamiltonian would have to be replaced with sums over momentum
1335: states and Coulomb interactions between electrons and holes would have
1336: to be included. Such a model without photons was studied by Keldysh
1337: and Kopaev \cite{keldysh1} and for coherently driven semiconductor by
1338: Schmitt-Rink et al. \cite{schmitt}. If we treat the Coulomb
1339: interaction on the mean-field level then the natural extension of the
1340: Keldysh mean-field wave-function to account for photons will be
1341: %
1342: \begin{equation} 
1343: \label{varwfn2} 
1344: |\Psi_0> = e^{\lambda 
1345: \psi_0^\dagger}\prod_{\vec k} [ u_{\vec k} + v_{\vec k} b^\dagger_{k}
1346: a_{k} ]|0\rangle \;\;\;,  
1347: \end{equation}
1348: %
1349: which is of analogous form to that for the two-level systems case
1350: (\ref{varwfn3}). Then the influence of the various decoherence
1351: processes can be included in exactly the same way as described in
1352: Section \ref{model}.
1353: 
1354: At the level of mean-filed, we do not expect that the change of the 
1355: model from that of two-level oscillators to free electron and holes with a 
1356: Coulomb interaction would cause any dramatic differences. It was
1357: shown within the mean-field techniques of Keldysh
1358: \cite{keldysh1,keldysh2,nozieresex1} that the Coulomb interaction has
1359: a pairing effect and leads to a formation of a coherent excitonic
1360: insulator phase and this would only be enhanced by the dipole
1361: coupling.
1362: 
1363: At high excitation densities (large photon fields) the dipole
1364: interaction between excitons and photons is the dominant
1365: interaction. In the case of a driven excitonic system Schmitt-Rink
1366: {\it at al} \cite{schmitt} pointed out that for a very large pumping,
1367: and thus high excitation densities, the Coulomb interaction is just a
1368: very minor correction to the dominant dipole coupling. They also state
1369: that, in the absence of the Coulomb interaction, their results for
1370: propagating electrons and holes are equivalent to those obtained for
1371: an ensemble of independent two-level oscillators as optical
1372: transitions with different {\bf k} decouple.
1373: 
1374: The Coulomb interaction would be more important at low densities.
1375: However, in this case, the dominant contribution arising from the 
1376: interaction between an electron and hole within the same
1377: exciton, is taken into account in model studied in this work. All
1378: other Coulomb interactions are much weaker and, as with the dipole
1379: coupling, favour condensation. 
1380: 
1381: Although we do not expect qualitative differences, it would be
1382: interesting, especially at low excitation densities, to perform similar
1383: calculations in the momentum representation with the Coulomb
1384: interaction. This would allow one to study the relative influence of the
1385: two independent pairing mechanisms, namely the dipole coupling and
1386: Coulomb interactions, on the condensate.
1387: 
1388: Summarising, our method of including and studying the decoherence
1389: effects is general and can be applied in the same way as in this work
1390: to propagating excitons as well as to coherently driven
1391: condensates. The coherent photon field in the driven system is an
1392: external, fixed, parameter and not a self-consistent field satisfying
1393: equation (\ref{saddlepointF2}). In contrast the gap in the density of
1394: states, proportional to the coherent field amplitude, is present
1395: exactly as for an equilibrium condensate.  The model can be applied to
1396: obtain a qualitative description of the physical behaviour even for
1397: propagating or weakly bound excitons. It would give similar
1398: predictions to the model based on propagating electrons and holes with
1399: Coulomb interactions treated within the mean-field approximation. It
1400: does not, however, include screening and other non mean-fields
1401: effects.
1402: 
1403: If the microscopic origins of the decoherence processes are known for
1404: a particular experimentally studied system, the theory could be
1405: extended to include this information. The detailed account of the
1406: coupling constants and the density of states for the environment can
1407: be easily included within this framework. In this work we have assumed
1408: a static bath or Markovian (Ohmic) bath but in general the bath
1409: propagators are frequency dependent and thus $\langle \psi \rangle$
1410: would depend on frequency as in the Eliashberg theory of gapless
1411: superconductors \cite{eliash,scalapino}. The phenomenological
1412: constants $\gamma_1$ and $\gamma_2$ can thus be, in principle,
1413: obtained from this microscopic calculation for a particular system.
1414: 
1415: The results presented in this work are performed at zero temperature
1416: for which the summations over the Matsubara frequencies in
1417: (\ref{saddlepointF2}), (\ref{pol}) and (\ref{inv}) become integrals
1418: and are performed numerically. The extension to finite temperatures is
1419: straightforward. At finite temperatures the summations over a discrete
1420: $\nu_n$ can be performed using the standard techniques
1421: \cite{green} and thus finite temperatures could be easily studied.
1422: 
1423: A much more important extension than the finite temperature case is
1424: the problem of non-equilibrium systems. For a system with very strong
1425: pumping and decay processes the thermal distribution of the relevant
1426: quasiparticles cannot be assumed. The influence of non-equilibrium
1427: distributions on the polariton condensate will be published elsewhere.
1428: 
1429: \section{Remarks on Experimental Realisation of BEC of polaritons}
1430: \label{exp}
1431: 
1432: While stimulated scattering into the lower polariton branch (reminiscent
1433: of Bose statistics) was clearly observed and reported by several groups
1434: \cite{dang,senbloch,boef,alex,bulkboser,deng,baumberg,angleboser,angle-cw-stimscat,nature},
1435: there is still no evidence of excitonic coherence, and therefore BEC of
1436: polaritons. The experiments that have been performed fall into two 
1437: categories. In the first, 
1438: polaritons are pumped coherently, with a laser having an angle of
1439: incidence chosen at a ``magic angle'' close to the bottom of the lower
1440: polariton branch. In the second set of experiments the pumping is
1441: incoherent. The laser pumping is tuned to be well above the ground
1442: state so that polaritons have to undergo several inelastic scattering
1443: events before reaching the ground state and thus loosing the initial
1444: coherence of the pump.
1445: 
1446: Although not yet measured, the potential coherence of polaritons in the
1447: experiments of the first category would be inherited from the pump and
1448: the behaviour could be explained in terms of parametric
1449: oscillation. Experiments of the second category could be potential
1450: candidates for spontaneous BEC of polaritons but evidence for it are
1451: very sparse. The coherence of the photon field
1452: emitted at polariton frequency has recently been measured in one
1453: experiment \cite{deng} but this alone does not prove that there is a
1454: coherence in the excitonic part. A laser is an example where coherent
1455: photon field is generated without large excitonic coherence. A more
1456: direct evidence of excitonic coherence is necessary. Large excitonic
1457: coherence would map to a large gap in the excitation spectrum which
1458: should be seen in the the incoherent luminescence.
1459: 
1460: Experiments on cavity polaritons which report the stimulated
1461: scattering into the lower polariton branch are performed at low
1462: densities to avoid the fermionic phase space filling effect, so that
1463: two clear polariton peaks could be seen.  For such densities the gap
1464: in the excitation spectrum would be very small and easily suppressed
1465: by the decoherence processes in the sample. Indeed such a gap has
1466: never been observed in the photoluminescence spectrum. The attempt to
1467: increase the density of polaritons in these experiments results not in
1468: the formation of the condensate but in a switching into the
1469: weak-coupling regime and lasing.
1470: 
1471: This is not surprising since the increase in the density of polaritons
1472: in these experiments is obtained by increasing the pumping of
1473: excitons. Incoherent pumping is a pair-breaking decoherence process
1474: and thus the increase in the pumping intensity results in the increase
1475: of not only the density of excitons but also decoherent scattering. As
1476: shown by Eastham and Littlewood \cite{paul}, the fermionic structure
1477: of excitons does not prevent condensation, even at very high
1478: densities. Thus, in the current experiments, we suggest that it is not
1479: a phase-space filling effect which lead to a laser as the pump
1480: intensity is increased, but the increase in the decoherence
1481: strength. If the polariton condensate is ever to be observed an
1482: increase in density without an increase in decoherence is necessary.
1483: 
1484: Localised and tightly bound excitons, such as excitons in disordered
1485: quantum wells or Frenkel-type excitons in organic compands, seem to be
1486: better candidates for observing a polariton condensate than the
1487: high-quality GaAs quantum wells with weakly bound, delocalised
1488: excitons. Static disorder is not a pair-breaking effect and would have
1489: a weaker influence on the condensate than screening and ionisation in
1490: the case of delocalised or weakly bound excitons.  Additionally the
1491: polariton splitting reported in organic materials is as large as 80
1492: meV \cite{organicpol1,organicpol2} in comparison to the upper bound of
1493: 20 meV in GaAs. Another very good candidate for observation of
1494: effects described in this work would be atoms in solid state, glass
1495: spheres, dilute atomic gases in which the dephasing is particularly
1496: weak in comparison to the dipole coupling.
1497: 
1498: \section{Summary}
1499: \label{sum}
1500: 
1501: We have studied the equilibrium Bose condensation of cavity polaritons
1502: in the presence of decoherence. It has been shown \cite{us0} that the
1503: widely used 
1504: Langevin equations with the constant, frequency independent decay
1505: rates are not valid for systems with large polarisation in which the
1506: gap in the density of states is present. In the regime of weak
1507: decoherence these processes have to be treated
1508: self-consistently, in such a way that the gap in the density of states is
1509: taken into account. We have proposed a self-consistent method
1510: analogous to the Abrikosov and Gor'kov theory of gapless
1511: superconductivity \cite{abrikosov-gorkov} which allow us to study all
1512: regimes of the system as the decoherence is changed from zero to large
1513: values and for a wide range of excitation densities.
1514: 
1515: We have found that at small decoherence the polariton condensate is
1516: protected by the energy gap in the excitation spectrum. The gap is
1517: proportional to the coherent field amplitude and thus the excitation
1518: density, so the condensate, unlike the excitonic insulator
1519: \cite{nozieresex1}, is more robust at high densities. This gap gets
1520: smaller and eventually is completely suppressed as the decoherence is
1521: increased.  We have shown that there is a regime, analogous to the
1522: gapless superconductor, when the coherent fields are present without
1523: an energy gap. This regime, at very high excitation densities, has
1524: most of the features of a photon laser.
1525: 
1526: We have studied the influence of two different types of processes,
1527: both pair-breaking and non-pair-breaking ones as well as inhomogeneous
1528: broadening of the energy levels. We have shown that only the
1529: pair-breaking processes can lead to phase transitions.
1530: Non-pair-breaking processes and inhomogeneous broadening of energies
1531: can quantitatively change the behaviour of the system but cannot
1532: prevent condensation.
1533: 
1534: We have studied a general form of the interactions, introducing
1535: pair-breaking and non-pair-breaking processes and thus our theory is
1536: applicable to a wide range of systems. For more detailed results, the
1537: particular origin of the interactions and thus the particular density
1538: of states for the baths have to be taken into account.
1539: 
1540: We have studied the whole phase diagram of the system given by the
1541: Hamiltonian (\ref{H}) for different values of the decoherence
1542: strength, inhomogeneous broadening and the excitation densities and
1543: established the crossover between an isolated polariton condensate and
1544: a photon laser as the decoherence strength is increased.
1545: 
1546: Our results suggest that, in contrast to traditional lasers, 
1547: coherent light can be generated without a population inversion. This
1548: work generalised the existing laser theories to include the gap in the
1549: excitation spectrum caused by the coherence in a media in the low
1550: decoherence regime.
1551: 
1552: \section{Acknowledgement}
1553: We would like to thanks F. M. Marchetti and P. R. Eastham for helpful
1554: discussions concerning mathematical techniques. 
1555: MHS acknowledge the support of research fellowship from Gonville and
1556: Caius College, Cambridge.
1557: 
1558: \begin{references}
1559: 
1560: \bibitem{abrikosov-gorkov}
1561: A.~A.~Abrikosov, L.~P.~Gor'kov, JETP {\bf 12}, 1243 (1960)
1562: 
1563: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1564: \bibitem{atompol} 
1565: M.~G. Raizen, R.~J. Thompson, R.~J. Breccha, H.~J. Kimble, and H.~J.  
1566: Carmichael, {\em Phys. Rev. Lett.} {\bf 63}, 240 (1989).
1567: 
1568: \bibitem{cavpol}
1569: C. Weisbuch, M. Nishioka, A.Ishikawa, and Y.Arakawa, {\em
1570: Phys. Rev. Lett.} {\bf 69}, 3314 (1992).
1571: 
1572: \bibitem{bulkcavpol} 
1573: A. Tedicucci, Y. Chen, V. Pellegrini, M. Borger, L. Sorba, F. Beltram,
1574: and F. Bassani, {\em Phys. Rev. Lett.} {\bf 75}, 3906 (1995).
1575: 
1576: \bibitem{organicpol1} 
1577: D.~G. Lidzey, D.~D.~C. Bradley, M.~S. Skolnick, T.~Virgili, S.~Walker,
1578: and D.~M. Whittaker, {\em Nature} {\bf 395}, 53 (1998).
1579: 
1580: \bibitem{organicpol2}
1581: D.~G. Lidzey, D.~D.~C. Bradley, T.~Virgili, A.~Armitage,
1582: M.~S. Skolnick, and S.~Walker, {\em Phys. Rev. Lett.} {\bf 82}, 3316 (1999).
1583: 
1584: \bibitem{chargedpol}
1585: R. Rapaport, R. Harel, E. Cohen, A. Ron, and E. Linder,
1586: {\em Phys. Rev. Lett.} {\bf 84}, 1607 (2000).
1587: 
1588: %%%%%%%%%%%%%%%%
1589: \bibitem{JJ}
1590: P. Barbara, A. B. Cawthorne, S. V. Shitov, and C. J. Lobb, 
1591: { \em Phys. Rev. Lett.} {\bf 82}, 1963 (1999).
1592: 
1593: \bibitem{paul} 
1594: P.~R. Eastham, P.~B. Littlewood, {\em Solid State Commun.} {\bf
1595: 116}, 357 (2000), { \em Phys. Rev. B} {\bf 64}, 235101 (2001). 
1596: 
1597: 
1598: \bibitem{laser} 
1599: For the quantum theory of the laser developed by
1600: Haken, Risken, Lax, Louisell, Scully and Lamb see M.~O. Scully and
1601: M.~S. Zubairy, {\em Quantum Optics} (Cambridge University Press,
1602: Cambridge, U.K., 1997).
1603: 
1604: 
1605: \bibitem{haken}
1606: H.~Haken, {\em Rev. Mod. Phys.} {\bf 47}, 67 (1975); 
1607: {\em Laser Theory}, Springer-Verlag 1984.
1608: 
1609: \bibitem{us0} 
1610: 
1611: M.~H.~Szymanska, P.~B. Littlewood, unpublished.  In
1612: (\ref{maxwell-bloch}) the life-time for a single two-level oscillator
1613: is generalised to many two-level systems without taking into account
1614: any collective effects. It can be shown that if we couple every
1615: two-level system to its own distinct bath we recover Equation
1616: (\ref{maxwell-bloch}). Such a method and so (\ref{maxwell-bloch}) is
1617: unphysical if the interactions between two-level system generated for
1618: example by the photon field is substantial. The environment needs to
1619: be coupled to the whole ensemble of two-level oscillators and averaged
1620: out in the selfconsistent way.
1621: 
1622: 
1623: \bibitem{wrong}
1624: K. Hannewald, S Glutsch and F. Bechstedt, {\em
1625: J. Phys. Condens. Matter} {\bf 13}, 275 (2001)  
1626: 
1627: \bibitem{keldysh1}
1628: L. V. Keldysh and Y. V. Kopaev, { \em Fiz. Tverd. Tela} {\bf 6}, 2791
1629: (1964), [Sov. Phys. Solid State {\bf 6}, 2219, (1965)].
1630: 
1631: \bibitem{keldysh2}
1632: L. V. Keldysh and A. N. Kozlov, {\em Zh. Eksp. Teor. Fiz.}  
1633: {\bf 54}, 978 (1968), [Sov. Phys. JETP {\bf 27}, 521 (1968)].
1634: 
1635: \bibitem{zittart}
1636: J.~Zittartz, {\em Phys. Rev.} {\bf 164}, 575 (1967)
1637: 
1638: \bibitem{us}
1639: M.~H.~Szymanska,  P.~B. Littlewood,
1640: { \em Solid State Commun.} {\bf 124}, 103 (2002)
1641: 
1642: \bibitem{schmitt}
1643: S. Schmitt-Rink, D.~S. Chemla, and H. Haug, { \em Phys. Rev. B} {\bf
1644: 37}, 941 (1988).
1645: 
1646: \bibitem{nozieresex1}
1647: C. Comte and P. Nozi{\`{e}}res, {\em J. Phys.} (Paris) {\bf 43}, 1069
1648: (1982).
1649: 
1650: \bibitem{eliash}
1651: G. M. Eliashberg, {\em Zh. Eksperim. i Teor. Fiz.} {\bf 38}, 966
1652: (1960), [Sov. Phys. JETP {\bf 11}, 696 (1960)].
1653: 
1654: \bibitem{scalapino}
1655: D.~J.~Scalapino, in {\em Superconductivity}, edited by R.~D. Parks,
1656: volume~1, chapter~10, pages 449--560, Marcel Dekker, Inc., New York,
1657: 1969.
1658: 
1659: \bibitem{green} 
1660: G. Rickayzen, {\em Green's Functions and Condensed Matter}
1661: (Academic Press Inc., London, 1987).
1662: 
1663: 
1664: \bibitem{dang}
1665: L.~S. Dang, D.~Heger, R.~Andr{\'{e}}, F.~B{\oe}uf, and R.~Romestain,
1666: { \em Phys. Rev. Lett.} {\bf 81}, 3920 (1998).
1667: 
1668: \bibitem{senbloch}
1669: P. Senellart and J. Bloch, { \em Phys. Rev. Lett.} {\bf 82},  1233  (1999).
1670: 
1671: \bibitem{boef} F. Boef, R. Andre, R. Romestain, L. S. Dang, {\em Phys. Rev. B}
1672: {\bf 62}, R2279 (2000).
1673: 
1674: \bibitem{alex} A. Alexandrou, G. Bianchi, E. Peronne, B. Halle, F. Boeuf,
1675: R. Andre, R. Romestain, L. S. Dang, {\em Phys. Rev. B}
1676: {\bf 64}, 233318 (2001).
1677: 
1678: \bibitem{bulkboser}
1679: V. Pellegrini, R. Colombelli, L. Sorba, and F. Beltram, { \em Phys. Rev. B}
1680: {\bf 59}, 10059 (1999).
1681: 
1682: \bibitem{deng} H. Deng, G. Weihs, C. Santori, J. Bloch, Y. Yamamoto, { \em
1683: Science} {\bf 298}, 199 (2002).
1684: 
1685: \bibitem{baumberg} 
1686: J. J. Baumberg, P. G. Savvidis, R. M. Stevenson, A. I. Tartakovskii,
1687: M. S. Skolnick, D. M. Whittaker, and J. S. Roberts, { \em
1688: Phys. Rev. B} {\bf 62}, R16247 (2000).
1689: 
1690: \bibitem{angleboser}
1691: P.~G. Savvidis, J.~J. Baumberg, R.~M. Stevenson, M.~S. Skolnick, D.~M.
1692: Whittaker, and J.~S. Roberts, { \em Phys. Rev. Lett.} {\bf 84}, 1547
1693: (2000).
1694: 
1695: \bibitem{angle-cw-stimscat} 
1696: R.~M. Stevenson, V.~N. Astratov, M.~S. Skolnick, D.~M. Whittaker,
1697: E.~Emam-Ismail, A.~I. Tartakovskii, P.~G. Savvidis, J.~J. Baumberg,
1698: and J.~S.  Roberts, {\em Phys. Rev. Lett.} {\bf 85}, 3680 (2000).
1699: 
1700: \bibitem{nature} 
1701: M. Saba, C. Ciuti, J. Bloch, V. Thierry-Mieg, R. Andre, Le Si Dang,
1702: S. Kundermann, A. Mura, G. Bongiovanni, J. L. Staehli, B. Deveaud,
1703: {\em Nature} {\bf 414}, 731 (2001)
1704: 
1705: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1706: \bibitem{atoms}
1707: C. Liu, Z. Dutton, C. H. Behroozi, L. V. Hau, {\em Nature} {\bf 409},
1708: 490 (2001) 
1709: 
1710: 
1711: 
1712: 
1713: \end{references}
1714: \end{multicols}
1715: 
1716: \end{document} 
1717:    
1718: 
1719: 
1720: 
1721: 
1722: 
1723: