1: \documentclass[preprint,prb,aps,showpacs]{revtex4}%
2: \usepackage{amsfonts}
3: \usepackage{amsmath}
4: \usepackage{amssymb}
5: \usepackage{graphics}
6: \usepackage{graphicx}%
7: \setcounter{MaxMatrixCols}{30}
8: %TCIDATA{OutputFilter=latex2.dll}
9: %TCIDATA{Version=5.00.0.2570}
10: %TCIDATA{LastRevised=Thursday, September 22, 2005 10:56:08}
11: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
12: %TCIDATA{<META NAME="SaveForMode" CONTENT="1">}
13: %TCIDATA{Language=American English}
14: \renewcommand{\textfraction}{0}
15: \renewcommand{\floatpagefraction}{0}
16: \renewcommand{\topfraction}{1}
17: \renewcommand{\bottomfraction}{1}
18: \setcounter{totalnumber}{2}
19: \setlength{\floatsep}{2cm}
20: \begin{document}
21: \title{Theory of emission from an active photonic lattice}
22: \author{Weng W. Chow}
23: \affiliation{{Sandia National Laboratories, Albuquerque, NM 87185-0601}}
24: \date{July 12, 2005}
25:
26: \pacs{}
27:
28: \begin{abstract}
29: The emission from a radiating source embedded in a photonic lattice is
30: calculated. The analysis considers the photonic lattice and free space as a
31: combined system. Furthermore, the radiating source and electromagnetic field
32: are quantized. Results show the deviation of the photonic lattice spectrum
33: from the blackbody distribution, with intracavity emission suppressed at
34: certain frequencies and enhanced at others. In the presence of rapid
35: population relaxation, where the photonic lattice and blackbody populations
36: are described by the same equilibrium distribution, it is found that the
37: enhancement does not result in output intensity exceeding that of the
38: blackbody at the same frequency. However, for slow population relaxation, the
39: photonic lattice population has a greater tendency to deviate from thermal
40: equilibrium, resulting in output intensities exceeding those of the blackbody,
41: even for identically pumped structures.
42:
43: \end{abstract}
44: \maketitle
45:
46:
47: \section{\bigskip Introduction}
48:
49: One of the many novel optical phenomena exhibited by photonic lattices is the
50: modification of spontaneous emission properties. \cite{pc1,pc2} A photonic
51: lattice can funnel radiation into narrow energy bands, where exceedingly high
52: intensities at photonic lattice bandedges have been predicted theoretically
53: and observed experimentally. \cite{piqeat,lin1,fleming,li} A question is
54: whether the peak intensities exceed those of a blackbody under similar
55: experimental conditions. \cite{dowling,lin,lin2,narayanaswamy,trupke,luo} The
56: answer is important for scientific understanding and can impact the
57: development of new light sources.
58:
59: It is generally agreed that the higher photonic-lattice density of states will
60: increase intracavity intensity. The debate concerns the output intensity in
61: comparison with that of a blackbody. Arriving at an answer is difficult
62: experimentally because it is difficult to ensure that the comparison is made
63: under similar conditions. Theoretically, the difficulty lies with the
64: treatment of the matter and optical aspects of the problem. \cite{siegman} The
65: derivation of the matter equations requires knowledge of the normal modes of
66: the optical structure, preferably in the form of an orthonormal basis.
67: However, such a basis set is not rigorously defined for a finite photonic
68: lattice with outcoupling loss. This problem occurs also in laser theory, where
69: one usually begins with the Fox-Li modes for a Fabry-Perot cavity with
70: perfectly reflecting mirrors, and introduces a loss mechanism to represent the
71: outcoupling. \cite{lamb} Such a phenomenological approach is inadequate for
72: the present problem because of the inconsistency arising from separating the
73: treatments of the eigenmode problem and the outcoupling effects.
74:
75: The approach taken in this paper\ considers the photonic lattice and free
76: space outside of the photonic lattice as one combined system (see Fig. 1). We
77: follow the method of an earlier paper on the linewidth of a Fabry-Periot laser
78: \cite{lang} in representing free space by a very large cavity. The photonic
79: lattice is approximated by a series of semitransparent interfaces. We begin
80: with discussing the one-dimensional geometry, \cite{dowling,chen} which we
81: will show to contain the essential features necessary for addressing our
82: question. Section II discusses the equations and boundary conditions obeyed by
83: the modes of our 'universe'. The determination of the eigenfrequencies and
84: eigenfunctions requires the simultaneous diagonalization of a usually large
85: matrix and the solution of a transcendental equation. A numerical procedure
86: for a photonic lattice of arbitrary size and interface transmission is
87: presented in Appendix A.
88:
89: In Sec. III, the radiation field is expanded in term of these large number of
90: modes and quantized. The radiation source is also treated quantum
91: mechanically, as an inhomogeneously broadened ensemble of two-level atoms
92: confined within the photonic lattice structure. The equations of motion for
93: the photon number and atomic populations are derived in this section. We
94: choose a fully quantized (i.e., quantized matter and field) treatment based on
95: Einstein's derivation of the Planck radiation law, which showed the importance
96: of a consistent treatment of stimulated and spontaneous emission processes.
97: \cite{einstein} Einstein was able to circumvent a fully quantized theory by
98: using Wien displacement law, which applied only to emission in free space. For
99: the photonic lattice, such a general relation does not exist.
100:
101: There are several recent calculations of photonic-lattice emission where the
102: emitting source is a classically described current. \cite{luo,chen} An
103: advantage of our treatment over these classical ones is that by paralleling
104: the Planck radiation law derivation, our comparison of photonic-lattice and
105: blackbody emission spectra is appreciably more straightforward. \ Moreover,
106: the fluctuation-dissipation theorem \cite{louisell} which is an essential
107: assumption in the classical calculations, \cite{luo,chen} appears as a result
108: in a fully quantized treatment because spontaneous emission is treated from
109: first principles.
110:
111: Section IV describes a spectrometer model used to determine the emission
112: spectrum. Section V uses the theory developed in the earlier sections to
113: investigate emission from an active photonic-lattice that is excited by an
114: external pump and allowed to equilibrate with a thermal bath via collisions.
115: The radiation field spectra measured inside and outside the photonic lattice
116: are described. Comparison of photonic lattice and blackbody emissions is
117: discussed for equilibrium and nonequilibrium situations.
118:
119: Section VI summarizes the extension to a 3-dimensional geometry. \ The
120: comparison between photonic lattice and blackbody spectra is made assuming a
121: spherically symmetric photonic-lattice dielectric function. The 3-d treatment
122: is important for three reasons. First, it proves that the theory can retrieve
123: Planck's blackbody distribution in the absence of a photonic lattice. Second,
124: it verifies the 1-d treatment in terms of containing the necessary physics for
125: answering the question of photonic lattice versus blackbody thermal emission.
126: Lastly, it points out the substantial increase in numerical demands with
127: increase dimensionality, thus justifying our concentration on the 1-d analysis
128: to facilitate physical understanding and tractability of numerics.
129:
130: \section{\bigskip Modes of the combined photonic-lattice and free-space
131: system}
132:
133: In this section, the eigenmodes for a photonic lattice coupled to the outside
134: world are derived using the model depicted in Fig. 1. The universe, which
135: embeds the photonic lattice, is represented by a very large cavity with
136: perfectly reflecting walls at $z=0$ and $z=L$. (End results are extrapolated
137: by taking the limit $L\rightarrow\infty$.) The photonic lattice is modeled as
138: a series of coupled resonators with semitransparent interfaces. Following
139: Spencer and Lamb, \cite{spencer} the semitransparent interfaces are described
140: as very thin surfaces with very large dielectric constants. As an
141: idealization, we use dielectric 'bumps' giving a dielectric permittivity%
142: \begin{equation}
143: \epsilon\left( z\right) =\epsilon_{0}\left[ 1+\frac{\eta}{\overline{k}}%
144: \sum\limits_{j=1}^{N_{pl}}\delta\left( z-z_{j}\right) \right] \label{1}%
145: \end{equation}
146: where $\eta=2\sqrt{\left( 1-T_{pl}\right) /T_{pl}}$, $T_{pl}$ is an
147: effective transmission at each interface located at $z_{j}$, $\overline{k}$ is
148: the average magnitude of the electromagnetic field wave vector and $N_{pl}$ is
149: the number of periods making up the photonic lattice. For brevity, we assume
150: the background permittivity inside the photonic lattice to be that of vacuum
151: $\epsilon_{0}$.
152:
153: Using the above dielectric function in Maxwell equations gives the following
154: differential equation for the eigenmodes of the combined photonic-lattice and
155: free-space system:%
156: \begin{equation}
157: \frac{d^{2}}{dz^{2}}u_{k}\left( z\right) =-\mu_{0}\epsilon\left( z\right)
158: \Omega_{k}^{2}u_{k}\left( z\right) \label{2}%
159: \end{equation}
160: where $\mu_{0}$ is the permeability in vacuum, $\Omega_{k}$ is the
161: eigenfrequency and $k$ labels the eigenmode. The boundary conditions are
162: obtained by first noting that the system is bounded by totally reflecting
163: surfaces, so that%
164: \begin{equation}
165: u_{k}(0)=u_{k}(L)=0 \label{3a}%
166: \end{equation}
167: Integrating Maxwell's equations across the bump gives the boundary conditions,%
168: \begin{align}
169: u_{k}(z_{j}^{+}) & =u_{k}(z_{j}^{-})\label{3b}\\
170: \frac{d}{dz}u_{k}\left( z_{j}^{+}\right) -\frac{d}{dz}u_{k}\left( z_{j}%
171: ^{-}\right) & =-\eta ku_{k}(z_{j})\ , \label{3c}%
172: \end{align}
173: where the superscripts $-$ and $+$ indicate the positions immediately before
174: and after an interface, respectively. Integrating by parts (\ref{2}) gives the
175: orthogonality relation%
176: \begin{equation}
177: \int_{0}^{L}dz\ \epsilon\left( z\right) u_{k}\left( z\right) u_{l}\left(
178: z\right) =\epsilon_{0}\delta_{k,l} \label{4}%
179: \end{equation}
180:
181:
182: Plotted in Fig. 2 are examples of eigenfunctions for a six-period ($N_{pl}=6$)
183: photonic lattice with effective interface transmission $T_{pl}=0.1$. Most of
184: the solutions are not resonant with the photonic lattice, so that mode
185: amplitude is negligible inside the photonic lattice, as shown in Fig. 2 (a).
186: Figures 2 (b) and 2 (c) show examples of photonic lattice modes, where the
187: latter figure clearly depicts the first derivative discontinuities at the interfaces.
188:
189: We show in Figs. 3 and 4 that the model can reproduce the photonic-lattice
190: properties relevant to our study. Figure 3 illustrates the formation of bands
191: and bandgaps, by plotting the frequencies of the photonic lattice modes [i.e.,
192: modes depicted Figs. 2 (b) and 2 (c)] versus the interface transmission. Not
193: plotted are the large number of free-space modes [Fig. 2 (a)], with mode
194: separation $\Omega=m\pi c/L\longrightarrow0$ as the system length
195: $L\longrightarrow\infty$. At $T_{pl}=0$, the photonic-lattice modes are simply
196: the modes of six uncoupled resonators, each of length $a$, i.e., they are
197: $N_{pl}$-fold degenerate and have frequencies $\Omega=m\pi c/a$ where $m$ is
198: an integer. The degeneracy is removed with coupling among sections of the
199: photonic lattice. The result is groupings of states separated by energy gaps,
200: as shown in the figure. As $N_{pl}$ become very large, the groups of states
201: become continuous bands, with the photonic-lattice modes residing entirely
202: within the shaded regions, and the free-space modes residing outside.\ At
203: $T_{pl}=1,$ the model (with a very long $L$) approximates the free-space situation.
204:
205: Figure 4 shows that the model can also reproduce the significant flattening of
206: the photonic-lattice dispersion at the bandedges. \ Plotted in the figure is
207: the dispersion for a twelve period ($N_{pl}=12$) photonic lattice, where the
208: points indicate the actual eigenfrequencies and the solid curve is a fit of
209: the data to illustrate the case of $N_{pl}\rightarrow\infty$. \ The flattening
210: of the dispersion at a bandedge results in a drastic increase in the density
211: of states. We define the density of states as $\rho(\omega)=dk_{0}/d\Omega$,
212: where following solid state convention $k_{0}$ is the wavevector with
213: vanishing interface reflectivity.\bigskip\ The effects of the large density of
214: states increase on intensity inside and outside of a photonic lattice is the
215: focus of this paper.
216:
217: \section{\bigskip Active medium and radiation field}
218:
219: To study the modification of emission characteristics by a photonic lattice,
220: we consider the situation of an ensemble of two-level atoms located inside a
221: photonic lattice. Each atom is labeled by $n$ and $j$, so that $\left\vert
222: a_{nj}\right\rangle $ and $\left\vert b_{nj}\right\rangle $ are the ground and
223: excited states, respectively, of an atom located at $z_{j}$ inside the
224: photonic lattice, with resonant energy $\hbar\omega_{n}$. Assuming the dipole
225: approximation, $z_{j}$ is a parameter locating the atom to a region that is
226: small compared to a wavelength, but large compared to the size of an atom. We
227: describe the radiation field emitted by these atoms in terms of the combined
228: system eigenmodes derived in the previous section, i.e.%
229:
230: \begin{equation}
231: E\left( z,t\right) =\sum_{k}\mathcal{E}_{k}\left[ a_{k}\left( t\right)
232: +a_{k}^{\dagger}\left( t\right) \right] \ u_{k}\left( z\right) \label{5}%
233: \end{equation}
234: where $\mathcal{E}_{k}=\sqrt{\hbar\Omega_{k}/\left( A\epsilon_{0}\right) }$,
235: $a_{k}^{\dagger}$ and $a_{k}$ are the photon creation and annihilation
236: operators, respectively, and $A$ is the cross section area of the structure.
237: From (\ref{5}), using Maxwell's equations and a dipole interaction, the
238: Hamiltonian for the matter and radiation-field system is \cite{Quantum
239: Optics,john1}
240: \begin{equation}
241: H=\sum_{n,j}\hbar\omega_{n}\left\vert b_{nj}\right\rangle \left\langle
242: b_{nj}\right\vert +\sum_{k}\hbar\Omega_{k}a_{k}^{\dagger}a_{k}-\sum
243: _{k,n,j}g_{kj}\left( \left\vert b_{nj}\right\rangle \left\langle
244: a_{nj}\right\vert a_{k}+a_{k}^{\dagger}\left\vert a_{nj}\right\rangle
245: \left\langle b_{nj}\right\vert \right) \ , \label{6}%
246: \end{equation}
247: where $g_{kj}=\mu\mathcal{E}_{k}u_{k}\left( z_{j}\right) $ and $\mu$ is the
248: dipole matrix element. Introducing the operators for the microscopic
249: polarization amplitude $p_{njk}\equiv\left\vert b_{nj}\right\rangle
250: \left\langle a_{nj}\right\vert a_{k}\exp\left[ -i\left( \omega_{n}%
251: -\Omega_{k}\right) t\right] $, the excited and ground state populations,
252: $\sigma_{anj}\equiv\left\vert a_{nj}\right\rangle \left\langle a_{nj}%
253: \right\vert $ and $\sigma_{bnj}\equiv\left\vert b_{nj}\right\rangle
254: \left\langle b_{nj}\right\vert $, respectively, and working in the Heisenberg
255: picture, \cite{louisell} we derive the equations of motion%
256:
257: \begin{align}
258: \frac{dp_{njk}}{dt} & =\frac{i}{\hbar}e^{-i\left( \omega_{n}-\Omega
259: _{k}\right) t}\sum_{k^{\prime}}g_{k^{\prime}j}\left( \sigma_{bnj}%
260: a_{k}a_{k^{\prime}}^{\dagger}-a_{k^{\prime}}^{\dagger}a_{k}\sigma
261: _{anj}\right) \ \label{7}\\
262: \frac{d\sigma_{anj}}{dt} & =\frac{i}{\hbar}\sum_{k}g_{kj}\left[
263: p_{njk}^{\dag}e^{-i\left( \omega_{n}-\Omega_{k}\right) t}-p_{njk}e^{i\left(
264: \omega_{n}-\Omega_{k}\right) t}\right] \label{8}\\
265: \frac{d\sigma_{bnj}}{dt} & =-\frac{i}{\hbar}\sum_{k}g_{kj}\left[
266: p_{njk}^{\dag}e^{-i\left( \omega_{j}-\Omega_{k}\right) t}-p_{njk}e^{i\left(
267: \omega_{j}-\Omega_{k}\right) t}\right] \label{9}%
268: \end{align}
269: Additionally, the photon number operator obeys,
270: \begin{equation}
271: \frac{da_{k}^{\dagger}a_{k}}{dt}=\frac{i}{\hbar}\sum_{n,j}g_{kj}\left[
272: p_{njk}^{\dag}e^{-i\left( \omega_{j}-\Omega_{k}\right) t}-p_{njk}e^{i\left(
273: \omega_{j}-\Omega_{k}\right) t}\right] \ . \label{10}%
274: \end{equation}
275: Assuming that the polarization decays because of dephasing collisions and that
276: the effective decay rate $\gamma$ is much larger than the rate of changes in
277: the active medium and photon populations, we can adiabatically eliminate the
278: polarization equation. Then, introducing the expectation values%
279:
280: \begin{align}
281: N_{k} & =\left\langle a_{k}^{\dagger}a_{k}\right\rangle \label{11}\\
282: N_{an} & =\sum_{j=1}^{N}\left\langle \sigma_{anj}\right\rangle \label{12}\\
283: N_{bn} & =\sum_{j=1}^{N}\left\langle \sigma_{bnj}\right\rangle \label{13}%
284: \end{align}
285: we obtain the working equations for our analysis:%
286:
287: \begin{align}
288: \frac{dN_{an}}{dt} & =\frac{2\mu^{2}}{\hbar\epsilon_{0}AL_{c}\gamma}%
289: \sum\limits_{k}\Omega_{k}\ \Gamma_{k}\ \left[ \left( N_{bn}-N_{an}\right)
290: N_{k}+N_{bn}\right] \ L\left( \omega_{n}-\Omega_{k}\right) \nonumber\\
291: & -\gamma_{r}\left[ N_{an}-f_{a}\left( \omega_{n},T\right) \right]
292: -\Lambda\left( \omega_{n}\right) N_{an}\label{14}\\
293: \frac{dN_{bn}}{dt} & =-\frac{2\mu^{2}}{\hbar\epsilon_{0}AL_{c}\gamma}%
294: \sum\limits_{k}\Omega_{k}\ \Gamma_{k}\ \left[ \left( N_{bn}-N_{an}\right)
295: N_{k}+N_{bn}\right] \ L\left( \omega_{n}-\Omega_{k}\right) \nonumber\\
296: & -\gamma_{r}\left[ N_{bn}-f_{b}\left( \omega_{n},T\right) \right]
297: +\Lambda\left( \omega_{n}\right) N_{an}\label{15}\\
298: \frac{dN_{k}}{dt} & =\frac{2\mu^{2}}{\hbar\epsilon_{0}AL_{c}\gamma}\sum
299: _{n}\Omega_{k}\ \Gamma_{k}\ \left[ \left( N_{bn}-N_{an}\right) N_{k}%
300: +N_{bn}\right] L\left( \omega_{n}-\Omega_{k}\right) -\gamma_{c}N_{k}
301: \label{16}%
302: \end{align}
303: where $N$ is the number of atoms, $L\left( x\right) =\left[ 1+\left(
304: x/\gamma\right) ^{2}\right] $ and%
305: \begin{equation}
306: \Gamma_{k}=\int_{0}^{L_{c}}dz\ \left\vert u_{k}\left( z\right) \right\vert
307: ^{2} \label{17}%
308: \end{equation}
309: is the mode confinement factor. In (\ref{14}) - (\ref{16}), the pump and decay
310: contributions are included phenomenologically, $\gamma_{c}$ is the photon
311: decay rate, $\Lambda\left( \omega_{n}\right) =\Lambda_{0}\exp\left[
312: \hbar\left( \omega_{0}-\omega_{n}\right) /k_{B}T_{p}\right] $ is the pump
313: rate, $\hbar\omega_{0}$ is the material bandgap energy and $\gamma_{r}$ is an
314: effective rate for the actual populations $N_{an}$ and $N_{bn}$ to relax to
315: the equilibrium distributions%
316:
317: \begin{align}
318: f_{a}(\omega_{n},T) & =Z_{n}\label{18}\\
319: \text{\ }f_{b}(\omega_{n},T) & =Z_{n}\exp%
320: %TCIMACRO{\QOVERD{(}{)}{-\hbar\omega_{n}}{k_{B}T}}%
321: %BeginExpansion
322: \genfrac{(}{)}{}{}{-\hbar\omega_{n}}{k_{B}T}%
323: %EndExpansion
324: \ , \label{19}%
325: \end{align}
326: where%
327: \begin{equation}
328: Z_{n}=\left[ 1+\exp%
329: %TCIMACRO{\QOVERD{(}{)}{-\hbar\omega_{n}}{k_{B}T}}%
330: %BeginExpansion
331: \genfrac{(}{)}{}{}{-\hbar\omega_{n}}{k_{B}T}%
332: %EndExpansion
333: \right] ^{-1}\ , \label{21}%
334: \end{equation}
335: $T_{p}$ and $T$ \ are the pump and reservoir temperatures. \ In our study,
336: (\ref{14}) to (\ref{16}) are solved numerically.
337:
338: \section{\bigskip Detector}
339:
340: To determine the spectra of the intracavity and output radiation, we use the
341: simple spectrometer model shown in Figure 5. In this model, two-level atoms
342: are placed in the region of interest. These atoms are prepared with only the
343: ground state $\left\vert a_{n}^{d}\right\rangle $ populated when the radiation
344: field is absent (zero detector temperature). The label $n$ indicates that the
345: level spacing between $\left\vert b_{n}^{d}\right\rangle $ and $\left\vert
346: a_{n}^{d}\right\rangle $ is $\omega_{n}^{d}$. The atoms interact weakly with
347: the radiation field to be measured, which excites some fraction of the atoms
348: to an excited state $\left\vert b_{n}^{d}\right\rangle $ that has some finite
349: lifetime $\gamma_{d}^{-1}$. Assuming a sufficiently fast detector response so
350: that the detector populations adiabatically follow the variations in the
351: photon number, the population in state $\left\vert b_{n}^{d}\right\rangle $
352: gives a measure of the radiation intensity ($\propto N_{k}$) in the region
353: occupied by the detector atom. The steady state upper detector state
354: population is%
355: \begin{equation}
356: N_{b}^{d}\left( \omega_{n}^{d}\right) =D\sum\limits_{k}\Omega_{k}N_{k}%
357: \frac{\gamma_{d}}{\gamma_{d}^{2}+\left( \omega_{n}^{d}-\Omega_{k}\right)
358: ^{2}}\int_{z_{d}}^{z_{d}+L_{d}}dz\ \left\vert u_{k}\left( z\right)
359: \right\vert ^{2} \label{22}%
360: \end{equation}
361: where $D=2\mu_{d}N_{d}/\left( \hbar\epsilon_{0}AL_{d}\gamma_{d}\right) $,
362: $\mu_{d}$ is the dipole matrix element between states $\left\vert b_{n}%
363: ^{d}\right\rangle $ and $\left\vert a_{n}^{d}\right\rangle $, $L_{d}$ is the
364: length of the detected region, and $N_{d}$ is the number of detector atoms.
365: Measuring this population for atoms of different $\omega_{n}^{d}$ gives the
366: spectrum within the region $z_{d}\leq z\leq z_{d}+L_{d}$. In this model,
367: $N_{d}$ and the decay rate $\gamma_{b}$ should be sufficiently large to
368: prevent saturation of the detector. \ On the other hand, too large a
369: $\gamma_{d}$ degrades spectral resolution. Alternately, one may use two level
370: atoms injected into the region of interest, and removed after a short time.
371: \cite{scully2}
372:
373: \section{Photonic lattice emission}
374:
375: We consider a twelve-period photonic lattice with $L_{c}=120\mu m$, $L=1.2cm$
376: and interface transmission $T_{pl}=0.01,$ $0.1$ and $0.4$. The eigenmodes are
377: determined by solving (\ref{2}) with the boundary conditions (\ref{3a}) -
378: (\ref{3c}). The results are used in (\ref{14}) - (\ref{16}), which are solved
379: numerically with a fourth-order Runge-Kutta finite difference method. The
380: input parameters are $\gamma=10^{12}s^{-1}$, $\gamma_{c}=10^{9}s^{-1}$,
381: $\Lambda_{0}=10^{10}s^{-1}$, $\omega_{0}=1.6\times10^{14}s^{-1}$, $\mu
382: =e\times1.3nm$, $N=601$ and $T_{p}=T=400K$.
383:
384: \subsection{Equilibrium}
385:
386: To relate to earlier studies, \cite{dowling,narayanaswamy,trupke,luo} we first
387: compare photonic lattice and blackbody emissions under thermal equilibrium
388: conditions. To do so, we perform the calculations for a rapid population
389: relaxation rate of $\gamma_{r}=10^{13}s^{-1}$, which ensures (verified after
390: the time integration) that the steady-state active-medium populations $N_{an}$
391: and $N_{bn}$ are to a good approximation given by the equilibrium
392: distributions $f_{a}(\omega_{n},T)$ and $f_{b}(\omega_{n},T)$, respectively.
393: The solid curves in Fig. 6 (a) show the calculated intracavity emission
394: spectra for three interface transmissions. \ In the figure, we define an
395: intracavity detector signal,%
396: \begin{equation}
397: S_{in}\left( \omega\right) \equiv\frac{N_{b}^{d}\left( \omega\right) }%
398: {D}=\sum\limits_{k}\Omega_{k}N_{k}\frac{\gamma_{d}}{\gamma_{d}^{2}+\left(
399: \omega-\Omega_{k}\right) ^{2}}\int_{0}^{L_{c}}dz\ \left\vert u_{k}\left(
400: z\right) \right\vert ^{2} \label{23}%
401: \end{equation}
402: where $N_{b}^{d}\left( \omega\right) $ is calculated using the steady-state
403: solution for $N_{k}$ in (\ref{22}). The figure shows two bands of
404: photonic-lattice states, where the frequency extent of the bands depends on
405: the interface transmission. Between the two bands is a photonic bandgap where
406: emission is strongly suppressed. By repeating the calculation with $T_{pl}=1$,
407: we obtain the corresponding blackbody spectrum (dashed curve). Comparison of
408: the curves clearly indicates the significant intensity enhancement inside a
409: photonic lattice, especially at the bandedges for $T_{pl}=0.01$.
410:
411: To determine the output spectrum, we place the spectrometer in the free-space
412: region. Figure 7 shows the output detector signal,
413: \begin{equation}
414: S_{out}\left( \omega\right) \equiv\frac{N_{b}^{d}\left( \omega\right) }%
415: {D}=\sum\limits_{k}\Omega_{k}N_{k}\frac{\gamma_{d}}{\gamma_{d}^{2}+\left(
416: \omega-\Omega_{k}\right) ^{2}}\int_{L-L_{c}}^{L}dz\ \left\vert u_{k}\left(
417: z\right) \right\vert ^{2} \label{24}%
418: \end{equation}
419: for the same interface transmissions as in Fig. 6. In contrast to inside the
420: photonic lattice, where there is significant optical intensity enhancement,
421: Fig. 7 indicates that the intracavity emission peaks are appreciably depressed
422: outside the photonic lattice. The strong intracavity enhancement by the
423: photonic-lattice density of states appears to be cancelled by an outcoupling
424: attenuation. This leaves the photonic lattice emission peaks to be essentially
425: independent of interface transmission. More importantly, these peaks lie at or
426: slightly below the blackbody emission curve.
427:
428: Simulations performed over a wide range of input parameters point towards the
429: result that as long as the active-medium populations $N_{an}$ and $N_{bn}$ are
430: in thermal equilibrium, the photonic-lattice output is always below that of
431: the blackbody. For instance, the spectra are insensitive to the choice of
432: $\gamma$. The blackbody spectrum is also insensitive to $\gamma_{d}$ because
433: of the weak frequency dependence of the blackbody photon density. However, for
434: the photonic-lattice, too large a $\gamma_{d}$ degrades the spectrometer
435: resolution and leads to lower spectral peaks. For the opposite situation, too
436: small a $\gamma_{d}$ introduces noise in the spectrum because of the
437: inadequate resolution of the system normal modes (i.e. because $L$ is
438: insufficiently large).
439:
440: \subsection{Nonequilibrium}
441:
442: To study active photonic-lattice operation in greater generality, we allow the
443: active medium populations to deviate from thermal equilibrium. The
444: investigation is performed by repeating the earlier calculations, keeping all
445: input parameters except $\gamma_{r}$ the same. Figure 8 illustrates the
446: changes in the excited-state population distribution $N_{bn}$, at steady state
447: and for decreasing population relaxation. When $\gamma_{r}$ is reduced to
448: $10^{11}s^{-1}$ from $10^{13}s^{-1}$, a slight difference emerges between the
449: excited-state populations of the identically pumped photonic-lattice and
450: blackbody active media. The solid curve in Fig. 8 (a) shows a noticeable
451: deformation of the photonic-lattice excited-state population distribution.
452: There is also a significant difference between $N_{bn}$ and $f_{b}\left(
453: \omega_{n},T\right) $ for $T=400K$ (dot-dashed curve), which are the actual
454: distribution and the asymptotic ($\gamma_{r}\rightarrow\infty$) equilibrium
455: distribution, respectively. Further reduction to $\gamma_{r}=10^{10}s^{-1}$
456: significantly increases the deviation of the photonic-lattice excited-state
457: population distribution from a Maxwell Boltzmann distribution [see solid
458: curve, Fig 8 (b)]. Holes are burned in the distribution because the population
459: relaxation is insufficiently fast to replenish the excited-state population
460: depleted by the spectrally relatively narrow radiation field emitted by
461: photonic lattice. There is also a change in the blackbody distribution [dashed
462: curve, Fig. 8(b)], to one that approximates a Maxwell Boltzmann distribution
463: at $T\approx500K$ (dotted curve). Since the active media in both structures
464: are identical, the difference between the photonic-lattice and blackbody
465: populations (solid and dashed curves, respectively) is from photonic-lattice effects.
466:
467: The effects of the population changes in Fig. 8 on the emission spectra are
468: depicted in Fig 9. Plotted on the y-axis is the relative emission intensity
469: inside (outside) the photonic lattice, which we define as $S_{in(out)}\left(
470: \omega\right) $ for the photonic lattice divided by $S\left( \omega\right)
471: $ for the blackbody. In spite of the large increase in excited state
472: population, we find that the intracavity and output relative intensities
473: remain basically unchanged when $\gamma_{r}$ is reduced from $10^{13}s^{-1}$
474: to $10^{11}s^{-1}$. In particular, the output photonic-lattice intensity
475: remains at or slightly below that of the blackbody (i.e., relative intensity
476: $\leq1$). However, the result changes considerably for $\gamma_{r}%
477: =10^{10}s^{-1}$. Here, the intensity within the photonic-lattice band
478: increases considerably relative to that of the blackbody both inside and
479: outside the cavity (solid curves). More importantly, the solid curve in Fig. 9
480: (b) clearly shows greater output intensity for the photonic lattice than the
481: blackbody throughout the emission band of the photonic lattice. This
482: enhancement of output emission occurs for identically pumped active regions
483: and is a result of a nonequilibrium population that shows significant hole
484: burning. The presence of nonequilibrium effects may be the cause for
485: experimental observations of metallic photonic-lattice emission exceeding that
486: of the blackbody. \cite{lin3} Note that the difference in output emission
487: spectra [solid and dashed curves in Fig. 9 (b)] comes from population
488: distributions that are, on the average, quite similar. That is, a
489: least-squares fit of the solid and dashed curves in Fig. 8 (b) will produce
490: Maxwell-Boltzmann's distributions that differ in temperature by less than 20K.
491: Therefore, measurement of average temperature will not identify the
492: experimental conditions leading to the photonic-lattice output emission
493: exceeding that of the blackbody. Rather, an energy-resolved measurement of the
494: emitter upper state population is necessary.
495:
496: \ The end results reached in our analyses involving equilibrium and
497: nonequilibrium situations are robust, i.e., they are relatively insensitive to
498: the choice of input parameters. While calculations performed with different
499: interface transmission show significant differences in spectral shapes and
500: intracavity intensities, the output intensities remain relatively constant
501: because of the mitigating influence of the coupling to free space.
502: Calculations are also performed for different number of photonic-lattice
503: periods. The results show negligible differences beyond $N_{pl}=10$, thus
504: verifying that the use of a 12-period photonic lattice does not lead to loss
505: of generality. Clearly noticeable are effects, such as differences in excess
506: output intensity with varying interface transmission, that are due to optical
507: nonlinearities in the nonequilibrium active medium. \ Such effects will not be
508: present in treatments using linear classical sources. \cite{luo}
509:
510: \section{ Extension to a 3-dimensional photonic lattice{}}
511:
512: \bigskip This section treats a 3-dimensional photonic lattice. Following
513: earlier quantum optical studies of photonics lattices, \cite{john} a
514: spherically symmetric dielectric function is assumed to simplify the numerics.
515: In spherical coordinates, the equation satisfied by the passive eigenmodes of
516: the combined photonic-lattice and free-space system, $u_{klm}\left(
517: r,\theta,\phi\right) $, is%
518: \begin{equation}
519: \frac{1}{r}\frac{\partial^{2}}{\partial r^{2}}\left( ru_{klm}\right)
520: +\frac{1}{r^{2}\sin\theta}\frac{\partial}{\partial\theta}\left( \sin
521: \theta\frac{\partial u_{klm}}{\partial\theta}\right) +\frac{1}{r^{2}\sin
522: ^{2}\theta}\frac{\partial^{2}u_{klm}}{\partial\phi^{2}}=-\mu_{0}%
523: \epsilon\left( r\right) \Omega_{klm}^{2}u_{klm}\ . \label{25}%
524: \end{equation}
525: Choosing the dielectric function%
526: \begin{equation}
527: \epsilon\left( r\right) =\epsilon_{0}\left[ 1+\frac{\eta}{\overline{k}}%
528: \sum\limits_{j=1}^{N_{pl}}\delta\left( r-r_{j}\right) \right] \ ,
529: \label{26}%
530: \end{equation}
531: \bigskip where $\eta$ and $\overline{k}$ are the same as in (\ref{1}), a
532: solution of (\ref{25}) between the photonic-lattice interfaces or in the
533: free-space region is%
534:
535: \begin{equation}
536: u_{klm}\left( r,\theta,\phi\right) =\left[ A_{kln}j_{l}\left( kr\right)
537: +B_{kln}\eta_{l}\left( kr\right) \right] Y_{lm}\left( \theta,\phi\right)
538: \ , \label{27}%
539: \end{equation}
540: where $j_{l}\left( \rho\right) $ and $\eta_{l}\left( \rho\right) $ are
541: spherical Bessel and Neumann functions, $Y_{lm}\left( \theta,\phi\right) $
542: is a spherical harmonic and the subscript $n$ indicates that the coefficients
543: $A_{kln}$ and $B_{kln}$ are for $r_{n}<r\leq r_{n+1}$. In order for a solution
544: to be finite at the origin and vanish at $r=r_{N_{pl}+1}$ (the end of the
545: region representing free space), we require
546: \begin{equation}
547: B_{kl1}=0 \label{28}%
548: \end{equation}
549: %
550:
551: \begin{equation}
552: A_{klN_{pl}+1}j_{l}\left( kr\right) +B_{klN_{pl}+1}\eta_{l}\left(
553: kr\right) =0\ . \label{29}%
554: \end{equation}
555: At the interfaces, the boundary conditions (\ref{3b}) and (\ref{3c}) demand%
556: \begin{equation}
557: A_{kln}j_{l}\left( kr\right) +B_{kln}\eta_{l}\left( kr\right)
558: -A_{kln+1}j_{l}\left( kr\right) -B_{kln+1}\eta_{l}\left( kr\right) =0
559: \label{30}%
560: \end{equation}%
561: \begin{gather}
562: A_{kln}\left[ -j_{l}^{\prime}\left( kr\right) +\eta j_{l}\left( kr\right)
563: \right] +B_{kln}\left[ -\eta_{l}^{\prime}\left( kr\right) +\eta\eta
564: _{l}\left( kr\right) \right] \nonumber\\
565: +A_{kln+1}j_{l}^{\prime}\left( kr\right) +B_{kln+1}\eta_{l}^{\prime}\left(
566: kr\right) =0 \label{31}%
567: \end{gather}
568: for $2\leq n\leq N_{pl}$ and $r=r_{1},r_{2,...,}r_{N_{pl}}$.
569:
570: The numerical solution is implemented similar to what is described in Appendix
571: A. A 6-period photonic lattice is considered, where the lattice constant is
572: $2\mu m$ and the interface transmission is $T_{pl}=0.05$. Coupled to the
573: photonic lattice is a 'free'-space region extending from $24\mu m<r\leq612\mu
574: m$. System dynamics is governed by the equations of motion (\ref{14}) -
575: (\ref{16}) with the photon-state index $k$ replaced by the three indices,
576: $k,l$ and $m$. Numerical analyses of the steady state solutions are performed
577: assuming the input parameters, $\gamma=2\times10^{12}s^{-1}$, $\gamma
578: _{r}=10^{13}s^{-1}$, $\gamma_{c}=10^{12}s^{-1}$, $\Lambda_{0}=10^{12}s^{-1}$,
579: $\omega_{0}=1.6\times10^{14}s^{-1}$, $\mu=e\times1.3nm$, $N=601$ and
580: $T_{p}=T=200K.$ Similar to the 1-dimensional case, the rates are chosen to
581: ensure reaching steady state with active-medium populations $N_{an}$ and
582: $N_{bn}$ described by Maxwell-Boltmann distributions.
583:
584: To obtain the intracavity emission spectrum, detector atoms are placed inside
585: the photonic-lattice structure. The probability of finding a photon with
586: frequency $\omega$ is proportional to%
587: \begin{equation}
588: S_{in}\left( \omega\right) =\sum_{klm}\Omega_{klm}N_{klm}\frac{\gamma_{d}%
589: }{\gamma_{d}^{2}+\left( \omega-\Omega_{klm}\right) ^{2}}\int_{0}^{2\pi}%
590: d\phi\int_{0}^{\pi}d\theta\int_{0}^{r_{Npl}}dr\ r^{2}\ \left\vert
591: u_{klm}\left( r,\theta,\phi\right) \right\vert ^{2}\ ,\label{32a}%
592: \end{equation}
593: so that the emission energy is proportional to $\hbar\omega S_{in}\left(
594: \omega\right) $. \ In Fig. 10, the solid curve is a plot of $\omega S_{in}$
595: as a function of frequency. It shows four narrow emission bands separated by
596: photonic bandgaps. Repeating the calculation using interface transmission
597: $T_{pl}=1$, gives the free-space emission spectrum (dashed curve). Examination
598: of the populations after steady state is reached verifies that both solid and
599: dashed curves are for identical Maxwell-Boltmann distributions at $T=200K$.
600: The curves clearly indicate intensity enhancement inside the photonic lattice.
601:
602: \bigskip To compare the output emission in a given direction, the detector
603: atoms are placed to give a signal,
604: \[
605: S_{out}\left( \omega\right) =\sum_{klm}\Omega_{klm}N_{klm}\frac{\gamma_{d}%
606: }{\gamma_{d}^{2}+\left( \omega-\Omega_{klm}\right) ^{2}}\int_{\phi_{d1}%
607: }^{\phi_{d2}}d\phi\int_{\theta_{_{d1}}}^{\theta_{_{d2}}}d\theta\int_{r_{d1}%
608: }^{r_{d2}}dr\ r^{2}\ \left\vert u_{klm}\left( r,\theta,\phi\right)
609: \right\vert ^{2}\ ,
610: \]
611: where $r_{d1}$, $r_{d2}$ are within the free-space region, $\theta_{d1}%
612: ,\theta_{d2}$ and $\phi_{d1},\phi_{d2}$ define the direction and collection
613: solid angle. The solid curve and dashed curves in Fig. 11 shows the output
614: photon-lattice and free-space emission spectra, respectively. These results
615: are obtained for $r_{d1}=100\mu m$, $r_{d2}=124\mu m$, $\phi_{d1}=\theta
616: _{d1}=0,$ $\phi_{d2}=2\pi$ and $\theta_{d2}=\pi/18$, which define emission
617: within a cone of $\pm10^{\circ}$ in the z-direction. Comparison of solid and
618: dashed curves indicates that, similar to the 1-dimensional case, peak
619: intensities measured outside the photonic-lattice structure do not exceed
620: those of the blackbody.
621:
622: \bigskip A test of our treatment is to see how well it reproduces Planck's
623: distribution for the frequency spectrum of free-space emission energy from a
624: thermal source. \ The dotted curve in Fig. 11 is proportional to $\omega
625: ^{3}\left[ \exp\left( \hbar\omega/k_{B}T\right) -1\right] ^{-1}$ with
626: temperature $T=200K$, and it depicts the shape of the blackbody frequency
627: spectrum according to Planck's formula. The agreement is good, considering
628: that there are several factors causing discrepancies. Two important ones are
629: truncating the optical modes\ at $l=13$ and limiting 'free'-space to $24\mu
630: m<r\leq612\mu m$, in order to maintain reasonable computation times. \ Even
631: so, over $10^{4}$ optical modes are used. Other factors contributing to the
632: differences include the presence of optical loss [$\gamma_{c}\neq0$ in
633: (\ref{16})], which is neglected in the derivation of Planck's distribution.
634: While increasing the number of optical modes improves agreement with Planck's
635: formula, it does not impact the blackbody versus photonic lattice emission
636: comparison. This is because the result of the photonic-lattice output
637: intensity spectrum being bounded inside the blackbody one applies separately
638: for each $l$.
639:
640: \section{ VI. Conclusion}
641:
642: In summary, the emission from an active photonic lattice is investigated using
643: a model consisting of an inhomogeneously broadened ensemble of two-level atoms
644: interacting with a multimode radiation field. A fully quantized (i.e.,
645: quantized atoms and quantized electromagnetic field) description is chosen to
646: provide a consistent description of stimulated and spontaneous emission.
647: Furthermore, to describe the modal properties of the radiation field of a
648: finite photonic lattice coupled to free space, the analysis considers the
649: photonic lattice and free space as one combined system. This circumvents a
650: long-standing inconsistency in quantum optics involving the decoupling of the
651: treatments of the cavity normal modes and outcoupling losses.
652:
653: Our approach gives the emission spectra for arbitrary photonic-lattice
654: configurations and reproduces Planck's blackbody radiation formula for thermal
655: emission in free space. Comparison of photonic-lattice and blackbody emission
656: shows appreciable modification of the blackbody spectrum by the photonic
657: lattice, where the redistribution of the photon density of states results in
658: suppression of radiation at certain wavelengths and enhancement at others. The
659: enhancement can give rise to high intracavity intensity peaks, especially at
660: the photonic-lattice bandedges. These intensity peaks are mitigated outside
661: the photonic lattice by the spectrally dependent outcoupling. For population
662: relaxation sufficiently fast to ensure the same equilibrium population
663: distribution in both structures, the photonic-lattice output intensity does
664: not exceed that of the blackbody at the same frequency. However, for slow
665: population relaxation, there is a greater tendency for a nonequilibrium
666: photonic-lattice population. Then, in the presence of population hole burning,
667: the intensity in certain regions of the photonic-lattice spectrum can exceed
668: that of the blackbody, even when both structures are identically pumped.
669:
670: \section{ V. Acknowledgments}
671:
672: This work was supported by the U. S. Department of Energy under contract No.
673: DE-AC04-94AL85000 and by a Senior Scientist Award from the Alexander von
674: Humboldt Foundation. \ The author thanks I. El-Kady, I. Waldmueller and S.
675: Wieczorek for helpful discussions.
676:
677: \section{ Appendix A: Numerical evaluation of system eigenmodes}
678:
679: In this appendix we describe a numerical procedure for evaluating the
680: eigenmodes of the combined photonic-lattice and free-space system. This
681: procedure applies for a photonic lattice of arbitrary size and interface
682: transmission. In region $n$, which may be any section of the photonic lattice
683: or the section representing free space, the solutions of (\ref{2}) have the form%
684:
685: \begin{equation}
686: u_{k}\left( z_{n}\right) =A_{k,n}\sin\left( kz_{n}\right) +B_{k,n}%
687: \cos\left( kz_{n}\right) \tag{A1}%
688: \end{equation}
689: where $k=\Omega_{k}/c$. Because of boundary condition (\ref{3a}),%
690: \begin{align}
691: A_{k,1} & \neq0\tag{A2}\\
692: B_{k,1} & =0 \tag{A3}%
693: \end{align}%
694: \begin{equation}
695: A_{k,N_{pl}+1}\sin\left( kz_{N_{pl}+1}\right) +B_{k,N_{pl}+1}\cos\left(
696: kz_{N_{pl}+1}\right) =0 \tag{A4}%
697: \end{equation}
698: From boundary conditions (\ref{3b}) and (\ref{3c}),%
699: \begin{align}
700: A_{k,n}\sin\left( kz_{n}\right) +B_{k,n}\cos\left( kz_{n}\right) &
701: =A_{k,n+1}\sin\left( kz_{n}\right) +B_{k,n+1}\cos\left( kz_{n}\right)
702: \tag{A5}\\
703: A_{k,n}\left[ -\cos\left( kz_{n}\right) +\eta\sin\left( kz_{n}\right)
704: \right] +B_{k,n}\left[ \sin\left( kz_{n}\right) +\eta\cos\left(
705: kz_{n}\right) \right] & =-A_{k,n+1}\cos\left( kz_{n}\right)
706: +B_{k,n+1}\sin\left( kz_{n}\right) \tag{A6}%
707: \end{align}
708: for $2\leq n\leq N_{pl}$.
709:
710: There are many approaches to numerically solve the above equations. We
711: describe below the one we followed in this paper. \ Basically, we look for the
712: values of $k$ satisfying (A4), where the mode amplitudes $A_{k,N_{pl}+1}$ and
713: $B_{k,N_{pl}+1}$ are obtained by solving a $2N_{pl}\times2N_{pl}$ matrix
714: equation
715: \begin{equation}
716: SU=D \tag{A7}%
717: \end{equation}
718: The matrix elements of $S$ are best defined by separating the even and odd
719: number rows. For $n=odd$,
720: \begin{align}
721: S_{i,i} & =-\sin\left( kz_{1}\right) \tag{A8}\\
722: S_{i,i+1} & =-\cos\left( kz_{i}\right) \tag{A9}%
723: \end{align}
724: and for $n=even$,
725: \begin{equation}
726: S_{i,i}=-\sin\left( kz_{i-1}\right) \tag{A10}%
727: \end{equation}
728: For $n\geq3$ and $n=odd$,%
729:
730: \begin{align}
731: S_{i,i-1} & =\cos\left( kz_{i}\right) \tag{A11}\\
732: S_{i,i-2} & =\sin\left( kz_{i}\right) \tag{A12}%
733: \end{align}
734: and for $n\geq4$ and $n=even$,%
735:
736: \begin{align}
737: S_{i,i-1} & =\cos\left( kz_{i-1}\right) \tag{A13}\\
738: S_{i,i-2} & =\sin\left( kz_{i-1}\right) +\eta\cos\left( kz_{i-1}\right)
739: \tag{A13}\\
740: S_{i,i-3} & =-\cos\left( kz_{i-1}\right) +\eta\sin\left( kz_{i-1}\right)
741: \tag{A14}%
742: \end{align}
743: All other matrix elements are zero. The elements of the column matrix $D$
744: vanishes except for%
745: \begin{align}
746: D_{1} & =-\sin\left( kz_{1}\right) \tag{A15}\\
747: D_{2} & =\cos\left( kz_{2}\right) -\eta\sin\left( kz_{2}\right)
748: \tag{A16}%
749: \end{align}
750:
751:
752: Equation (A7) is solved using the Gauss-Jordan method. In the solutions and
753: for $j=odd$, $U_{j}$ gives the coefficient $A_{k,j+1},$while $U_{j+1}$ gives
754: the coefficient $B_{k,j+1}$. At this stage, we have set $A_{k,1}=1$ so that
755: the eigenfunctions are unnormalized. \ We perform the normalization according
756: to (\ref{4}).
757:
758: \pagebreak
759:
760: \begin{thebibliography}{99} %
761:
762:
763: \bibitem {pc1}E. Yablonovitch, Phys. Rev. Lett. \textbf{58}, 2059 (1987).
764:
765: \bibitem {pc2}S. John, Phys. Rev. Lett. \textbf{58}, 2486 (1987).
766:
767: \bibitem {piqeat}P. Pigeat, D. Rouxel and B. Weber, Phys. Rev. B \textbf{57},
768: 9293 (1998).
769:
770: \bibitem {lin1}S.-Y. Lin, J. G. Fleming, E. Chow and J. Bur, Phys. Rev. B
771: \textbf{62}, R2243 (2000).
772:
773: \bibitem {fleming}J. G. Fleming, S. Y. Lin, I. El-Kady, R. Biswas and K. M.
774: Ho, Nature (London) \textbf{417}, 52 (2002).
775:
776: \bibitem {li}A.-Y. Li, Phys. Rev. B \textbf{66}, R241103 (2002).
777:
778: \bibitem {dowling}C. M. Cornelius and J. P. Dowling, Phys. Rev. A \textbf{59},
779: 4736 (1999).
780:
781: \bibitem {lin}S.-Y. Lin, J. Moreno and J. G. Fleming, Appl. Phys. Lett.
782: \textbf{83}, 380 (2003).
783:
784: \bibitem {lin2}S.-Y. Lin, J. G. Fleming and I. El-Kady, Opt. Lett.
785: \textbf{28}, 1909 (2003).
786:
787: \bibitem {narayanaswamy}A. Narayanaswamy and G. Chen, Phys. Rev. B.
788: \textbf{70}, 125101 (2004).
789:
790: \bibitem {trupke}T. Trupke, P. W\"{u}rfel and M. A. Green, Appl. Phys. Lett.
791: \textbf{84}, 1997 (2004).
792:
793: \bibitem {luo}C. Luo, A. Narayanaswamy, G. Chen and J. D. Joannopoulos, Phys.
794: Rev. Lett. \textbf{93}, 213905 (2004).
795:
796: \bibitem {siegman}For a text book discussion see Chap. 24, A. E. Siegman,
797: \textit{Lasers,} University Science Books, Mill Valley (1986).
798:
799: \bibitem {lamb}W. E. Lamb, Jr., Phys. Rev. \textbf{134}, A1429 (1964).
800:
801: \bibitem {lang}R. Lang, M. O. Scully and W. E. Lamb, Jr., Phys. Rev. A
802: \textbf{7}, 1788 (1973).
803:
804: \bibitem {chen}A. Narayanaswamy and G. Chen, Phys. Rev. B \textbf{70}, 125101 (2004).
805:
806: \bibitem {einstein}For a recent discussion, see: D. Kleppner, Physics Today,
807: February 2005, page30.
808:
809: \bibitem {louisell}W. H. Louisell, \textit{Quantum Statistical Properties of
810: Radiation,} Wiley, New York (1973).
811:
812: \bibitem {spencer}M. B. Spencer and W. E. Lamb, Jr., Phys. Rev. A \textbf{5},
813: 884 (1972).
814:
815: \bibitem {Quantum Optics}M. O. Scully and M. S. Zubairy, \textit{Quantum
816: Optics}, Cambridge University Press, Cambridge (1977).
817:
818: \bibitem {john1}N. Vats, S.John, K. Busch, Phys. Rev. A \textbf{65}, 043808 (2002).
819:
820: \bibitem {scully2}M. O. Scully and W. E. Lamb, Jr., Phys. Rev. A \textbf{166},
821: 246 (1968).
822:
823: \bibitem {lin3}S.-Y. Lin, J. Moreno and J. G. Fleming, Appl. Phys. Lett.
824: \textbf{84}, 1999 (2004).
825:
826: \bibitem {john}S. John and J. Wang, Phys. Rev. B \textbf{43}, 12772 (1991).
827:
828: \newpage
829: \end{thebibliography}
830:
831: \begin{center}
832: \bigskip\textbf{Figure Captions\\[2mm]}
833: \end{center}
834:
835: Fig. 1. Model of a photonic lattice connected to a large cavity approximating
836: the universe.
837:
838: Fig. 2. \ Eigenfunctions of a 6-period photonic-lattice coupled to free space
839: for interface transmission $T_{pl}=0.10$. The figure shows non-resonant (a)
840: and resonant (b and c) photonic-lattice modes.
841:
842: Fig. 3. Eigenfrequency versus interface transmission for a 6-period photonic
843: lattice. The points indicate the actual eigenmodes of the finite structure,
844: while the shaded regions illustrate the extent of the bands of
845: photonic-lattice states when the number of periods become very large.
846:
847: Fig. 4. Dispersion for a 12-period photonic lattice. The points are the
848: eigenmodes, the solid curve is a fit through these points and the dashed curve
849: shows the free-space dispersion.
850:
851: Fig. 5. Spectrometer model where the upper level decay prevents detector
852: saturation and approximates the drift of carriers to the electrodes in a
853: reverse-baised photodiode.
854:
855: Fig. 6. Photonic-lattice (solid curves) and blackbody (dashed curves)
856: intracavity emission spectra.
857:
858: Fig. 7. Photonic-lattice (solid curves) and blackbody (dashed curves) output
859: emission spectra.
860:
861: Fig. 8. Upper state population for photonic lattice (solid curve) and
862: blackbody (dashed curve) versus transition frequency for $\gamma_{r}=10^{11}$
863: (a) and $10^{10}s^{-1}$ (b). The dot-dashed curves show the equilibrium
864: distribution at $400K$. The dotted curve is the equilibrium distribution at
865: $500K$.
866:
867: Fig. 9. Relative intensity spectra inside (a) and outside (b) the photonic
868: lattice in Fig. 8. The curves are for $\gamma_{r}=10^{11}$ (dashed) and
869: $10^{10}s^{-1}$ (solid). Above the long-dashed line, the photonic-lattice
870: intensity is higher than the blackbody's.
871:
872: Fig. 10. 3-dimensional photonic-lattice and blackbody intracavity emission
873: spectra (solid and dashed curves, respectively).
874:
875: Fig. 11. Output emission spectra for 3-d photonic-lattice (solid curve),
876: blackbody (dashed curve) and Planck's distribution (dotted curve).
877: \end{document}