1: %% ****** Start of file template.aps ****** %
2: %%
3: %%
4: %% This file is part of the APS files in the REVTeX 4 distribution.
5: %% Version 4.0 of REVTeX, August 2001
6: %%
7: %%
8: %% Copyright (c) 2001 The American Physical Society.
9: %%
10: %% See the REVTeX 4 README file for restrictions and more information.
11: %%
12: %
13: % This is a template for producing manuscripts for use with REVTEX 4.0
14: % Copy this file to another name and then work on that file.
15: % That way, you always have this original template file to use.
16: %
17: % Group addresses by affiliation; use superscriptaddress for long
18: % author lists, or if there are many overlapping affiliations.
19: % For Phys. Rev. appearance, change preprint to twocolumn.
20: % Choose pra, prb, prc, prd, pre, prl, prstab, or rmp for journal
21: % Add 'draft' option to mark overfull boxes with black boxes
22: % Add 'showpacs' option to make PACS codes appear
23: % Add 'showkeys' option to make keywords appear
24: %\documentclass[aps,prl,preprint,superscriptaddress,showpacs,showkeys]{revtex4}
25: \documentclass[aps,prl,twocolumn,superscriptaddress,showpacs,showkeys]{revtex4}
26: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
27: %\documentclass[aps,prl,twocolumn,groupedaddress]{revtex4}draft
28: \usepackage{amsmath,amssymb,epsfig}
29: % You should use BibTeX and apsrev.bst for references
30: % Choosing a journal automatically selects the correct APS
31: % BibTeX style file (bst file), so only uncomment the line
32: % below if necessary.
33: %\bibliographystyle{apsrev}
34:
35: \begin{document}
36:
37: % Use the \preprint command to place your local institutional report
38: % number in the upper righthand corner of the title page in preprint mode.
39: % Multiple \preprint commands are allowed.
40: % Use the 'preprintnumbers' class option to override journal defaults
41: % to display numbers if necessary
42: \preprint{NSF-ITP-02-67}
43:
44: %Title of paper
45: \title{Real-time determination of free energy and losses in optical
46: absorbing media }
47:
48: % repeat the \author .. \affiliation etc. as needed
49: % \email, \thanks, \homepage, \altaffiliation all apply to the current
50: % author. Explanatory text should go in the []'s, actual e-mail
51: % address or url should go in the {}'s for \email and \homepage.
52: % Please use the appropriate macro foreach each type of information
53:
54: % \affiliation command applies to all authors since the last
55: % \affiliation command. The \affiliation command should follow the
56: % other information
57: % \affiliation can be followed by \email, \homepage, \thanks as well.
58: \author{C. Broadbent}
59: \affiliation{Department of Physics, Brigham Young University,
60: Provo, Utah, 84601}
61: \author{G. Hovhannisyan}
62: \affiliation{Department of Mathematics, Brigham Young University,
63: Provo, Utah 84601 }
64: \author{J. Peatross}
65: \affiliation{Department of Physics, Brigham Young University,
66: Provo, Utah, 84601}
67: \author{M. Clayton}
68: \affiliation{Department of Mathematics, Penn State University,
69: University Park, State College, Pennsylvania 16802}
70: \author{S. Glasgow}
71: %\email[]{Your e-mail address}
72: %\homepage[]{Your web page}
73: %\thanks{}
74: %\altaffiliation{}
75: \affiliation{Department of Mathematics, Brigham Young University,
76: Provo, Utah 84601 }
77:
78: %\email[]{Your e-mail address}
79: %\homepage[]{Your web page}
80: %\thanks{}
81: %\altaffiliation{}
82:
83:
84: %Collaboration name if desired (requires use of superscriptaddress
85: %option in \documentclass). \noaffiliation is required (may also be
86: %used with the \author command).
87: %\collaboration can be followed by \email, \homepage, \thanks as well.
88: %\collaboration{}
89: %\noaffiliation
90:
91: \date{\today}
92:
93: \begin{abstract}
94: % insert abstract here
95: We introduce notions of free energy and loss in linear, absorbing
96: dielectric media which are relevant to the regime in which the
97: macroscopic Maxwell equations are themselves relevant. As such we
98: solve a problem eluded to by Landau and Lifshitz \cite{LL84} in
99: 1958, and later considered explicitly by Barash and Ginzburg
100: \cite{BG76}, and Oughtsun and Sherman \cite{OS94}. As such we
101: provide physically-relevant real-time notions of "energy" and
102: "loss" in all analogous linear dissipative systems.
103:
104: \end{abstract}
105:
106: % insert suggested PACS numbers in braces on next line
107: \pacs{}
108: % insert suggested keywords - APS authors don't need to do this
109: %\keywords{}
110:
111: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
112: \maketitle
113:
114: % body of paper here - Use proper section commands
115: % References should be done using the \cite, \ref, and \label commands
116: %\section{\label{1} Introduction}
117: % Put \label in argument of \section for cross-referencing
118: %\section{\label{}}
119:
120: In a previous publication we showed that in many instances fast
121: and slow light \cite{GM70,CW82,HHDB99} is a manifestation of a
122: dielectric interacting differently with the early parts of an
123: electromagnetic pulse than with its later parts \cite{JOSAA}. For
124: example, slow light in a passive linear dielectric typically
125: corresponds to energy in the leading part of the pulse being
126: preferentially stored by the medium and then being largely
127: returned to the pulse's (backward) tail. Quantifying the extent to
128: which this process is possible was a primary motivation for this
129: work.
130:
131: Application of the principles presented in this letter will in
132: no-wise supersede the group velocity description of such
133: phenomena. The present development compliments other such analysis
134: by attaching precise notions of reversibility and irreversibility
135: to the medium's storage of the pulse energy. As such it also
136: solves a long outstanding problem regarding the real-time meaning
137: of "energy" and "loss" in linear dissipative systems.
138:
139: From Poynting's energy conservation theorem (Eq. \ref{Eq:Poyn})
140: the total energy density in a dielectric at time $t$, $u(t)$, is
141: the sum (Eq. \ref{Eq:utot}) of the field energy $u_{field}(t)$
142: (Eq. \ref{Eq:ufield}) and the medium-field interaction energy
143: $u_{int.}(t)$ (Eq. \ref{Eq:uint}):
144:
145: \begin{align}
146: \nabla \cdot {\bf S }+ \frac{\partial u}{\partial t} &= 0 \label{Eq:Poyn}\\
147: u(t) &= u_{field}(t)+u_{int.}(t) \label{Eq:utot}\\
148: u_{field}(t) &= \frac{E^2}{2} + \frac{H^2}{2}
149: \label{Eq:ufield}\\
150: u_{int.}(t) &= \int\limits_{-\infty}^t E(\tau)\dot{P}(\tau) \,
151: d\tau .\label{Eq:uint}
152: \end{align}
153:
154: Here we restrict to isotropic, temporally dispersive media in
155: which a scalar analysis suffices, and in which one may safely
156: suppress reference to the spatial coordinate ${\bf x}$. In
157: addition we restrict to linear media. Consequently we restrict to
158: the case in which the medium's response to an applied field ${\bf
159: E}$ is completely determined by a scalar, point-wise defined
160: susceptibility $\chi=\chi(\omega,{\bf x})$. For a basic discussion
161: of properties of the medium-field interaction energy $u_{int.}(t)$
162: in anisotropic media, see \cite{GWP01}. Since $u_{int.}(t)$
163: quantifies the net work the field has done against the polarized
164: medium in the course of creating the current system state, in the
165: following we will also call $u_{int.}(t)$ the (medium) internal
166: energy. In the sequel, notions of work done, and the ability to do
167: work in the future, will become pivotal in establishing an
168: unambiguous, dynamically relevant notion of energy allocation in
169: dielectric media.
170:
171: Landau and Lifshitz interpreted $u_{int.}(+\infty)$ as the energy
172: that is eventually lost to and dissipated by the medium over the
173: course of the medium-field interaction \cite{LL84}. This
174: asymptotic quantity only depends on the medium susceptibility
175: $\chi$ and the evolution of the electric field $E$, not on the
176: particulars of any microscopic model giving rise to $\chi$. This
177: suggests the possibility of establishing a real-time, model
178: independent notion of loss. Barash and Ginsburg considered this
179: question \cite{BG76}. They concluded, however, that a real-time
180: determination of losses is impossible without a microscopic model,
181: i.e. that a knowledge of the macroscopic susceptibility $\chi$ is
182: insufficient. Having concluded that a model independent notion of
183: loss is meaningless, they calculated (a certain notion of) loss
184: for specific models. In particular they generalized the work of
185: Loudon \cite{L70} concerning single Lorentz oscillator media by
186: making a straightforward extension of his notion of losses to
187: multiple Lorentz oscillators. Loudon's and, so, Barash and
188: Ginsburg's notion of losses amounts to identifying those terms in
189: (a certain expression of) the internal energy which explicitly
190: depend on phenomenological damping parameters. More convincingly,
191: their determination of the "energy" in the medium amounts to
192: summing the kinetic and potential energies of the individual
193: oscillators. We call this notion of energy the \emph{polarization}
194: energy, $u_{polarization}(t)$.
195:
196: We initially followed a certain extension of this reasoning. We
197: hoped to establish a unique map from an arbitrary susceptibility
198: $\chi(\omega)$ to an oscillator representation, which in turn
199: would establish a unique (polarization) energy, as well as losses
200: via Loudon's identification procedure. Unfortunately, we found
201: that generically a susceptibility can be mapped to many different
202: microscopic models. This property is perfectly analogous to the
203: one by which it is possible for distinct LRC circuits to yield
204: identical impedances. However, we then conjectured that all such
205: equivalent representations might yield identical \emph{values} for
206: the polarization energy. Unsurprisingly, one finds that this is
207: not the case dynamically. (Although asymptotically all such
208: representations must obviously agree:
209: $u_{polarization}(+\infty)=0$, and the losses are given, then, by
210: $u_{int}(+\infty)$, which, as mentioned, is completely determined
211: by $\chi$.)
212:
213: Figures \ref{fig:sprgms} and \ref{fig:free} demonstrate this
214: ambiguity. Given a double Lorentz oscillator susceptibility
215: $\chi(\omega) = \sum\nolimits_{n=1,2}
216: \omega_{p_n}^2/(\omega_n^2-i\gamma_n\omega-\omega^2)$ we examine
217: two different microscopic models for the same susceptibility. The
218: first model $\chi_{a}(\omega)$, is given by the explicit structure
219: of $\chi(\omega)$ as just written. It corresponds to two different
220: masses of equal charge on two different springs, each spring
221: having different restoring and damping parameters. The total
222: polarization is given by the net displacement of both charged
223: masses from their equilibrium positions:
224: $P=\textup{X}_{1_a}+\textup{X}_{2_a}=\chi_a E$. See
225: FIG.~\ref{fig:sprgms}(a).
226:
227: \begin{figure}
228: \includegraphics{SpringFigure.eps}
229: \caption{Macroscopic equivalent spring-mass systems for a double
230: Lorentz oscillator and their LRC circuit analogs. (a) The two
231: oscillator representation. (b) The LRC circuit analog of (a).
232: (c) The coupled oscillator representation. (d) The circuit analog of (c).\label{fig:sprgms}}
233: \end{figure}
234:
235: We use the tangency algorithm~\cite{GMCHB}, a transfer function
236: preserving algorithm intended for model reduction in LRC circuits,
237: to construct the second microscopic model $\chi_b(\omega)$.
238: Mechanically this model corresponds to coupled oscillators. See
239: FIG.~\ref{fig:sprgms}(c). The susceptibility for this microscopic
240: model is given in Eq. (\ref{Eq:chib}) (where we have mapped
241: $\omega$ to $i s$ for convenience). In this model the total
242: polarization is given by the displacement of the charged mass from
243: its equilibrium position $P=\textup{X}_{1_b}=\chi_b E$. It can be
244: shown that the coefficients $k_{n_b},\gamma_{n_b},m_{n_b}$ can be
245: found such that $\chi_a(\omega)=\chi_b(\omega)$ for all
246: frequencies $\omega$.
247: \begin{eqnarray}
248: \chi_b(s) = \frac{1}{k_{1_b}+\gamma_{1_b}s+m_{1_b}s
249: ^2+\frac{1}{\frac{1}{k_{2_b}+\gamma_{2_b}s}+\frac{1}{\gamma_{3_b}s}+\frac{1}{m_{3_b}s^2}}}
250: \label{Eq:chib}
251: \end{eqnarray}
252:
253: By inserting the two representations into the internal energy $u_{int}(t) =
254: \int\nolimits_{-\infty}^{t} E(\tau)\dot{P}(\tau)\,d\tau$ one can
255: find the kinetic and potential energies associated with each
256: representation. This polarization energy, the energy reckoned
257: instantaneously via the motion and position of the masses, will be
258: shown to be representation dependent. To demonstrate this, we
259: calculate the "microscopic" energetics associated with the
260: response of these macroscopically equivalent systems to an
261: impulse. That is we excite the two media, with susceptibility
262: models $\chi_a$ and $\chi_b$, via a delta function E-field at time
263: $t=0$, and then plot the losses for each representation. We assign
264: values to the parameters of the two oscillator, $\chi_a$
265: representation (the plasma frequencies, resonant frequencies, and
266: damping coefficients), and then find parameters for the coupled
267: oscillators', $\chi_b$ representation such that the
268: representations are macroscopically equivalent. Figure
269: \ref{fig:free} shows the evolution of the losses for the different
270: representations, these losses determined in the sense of Loudon,
271: Barash and Ginzburg. The corresponding differences between the
272: internal energy imparted to the media by the delta function
273: E-field (as shown by the piecewise constant curve in figure
274: \ref{fig:free}) and the plotted losses gives the polarization
275: energies for the two representations.
276:
277: \begin{figure}
278: \includegraphics{HeatFigure.eps}
279: \caption{The instantaneous losses of a
280: two oscillator (long dash), and a coupled oscillator (short dash) model of a
281: double Lorentz oscillator susceptibility
282: responding to an instigating delta function E-field at time
283: $t=0$. The interaction energy
284: $u_{int.}[E](t)$ (solid), and the irreversible energy $u_{irrev.}(t)$ (dot-dash) are also shown.\label{fig:free}}
285: \end{figure}
286:
287: The qualitative features of Figure \ref{fig:free} can be
288: understood intuitively: the delta excitation of the medium
289: instantaneously creates the polarization energies, which, then,
290: decrease as the systems dissipate these energies. However, as
291: clearly seen in the figures, the polarization energies for each
292: microscopic representation differ significantly. Thus the
293: polarization energy does in fact depend on the microscopic model
294: ascribed to the susceptibility $\chi(\omega)$. Of course, since
295: Barash and Ginzburg specified no underlying macroscopic physical
296: principle in their determination of "energy", this result is just
297: a consequence of their (lack of a) definition of energy.
298: Consequently, we will argue that a model-dependent (e.g.
299: polarization) energy is irrelevant, both from the point of view of
300: the relevant physical principles that should be required of a
301: viable macroscopic description, as well as from that of the
302: practical considerations of the energy storage and return process
303: mentioned at the beginning of this article.
304:
305: Before establishing relevant macroscopic principles of energy
306: allocation, we highlight the connection between mechanical and
307: electrical oscillators and the ambiguities associated with
308: representation in the context of the latter. The perfect analogy
309: between electrical LRC circuits and linear passive dielectric
310: media (where polarization $P$, susceptibility $\chi$ and E-field
311: $E$ , translate to charge $Q$, derivative of admittance
312: $\partial_t A$, and electromotive force $\mathcal{E}$,
313: respectively), allows one to reinterpret the polarization energy
314: evolutions implied in Fig. \ref{fig:free} as the evolutions of the
315: energies contained in the dispersive elements (inductors and
316: capacitors) of admittance-equivalent circuits. In the case of
317: electrical circuits, the plotted losses correspond to the losses
318: in the dissipative elements of the two circuits, i.e. the losses
319: through the resistors. This time we interpret the discrepancies in
320: these curves as indicating that different LRC circuits with the
321: same admittance can, temporarily, lose energy to their resistors
322: at different rates. Figures \ref{fig:sprgms}(b)and (d) give LRC
323: circuits corresponding to the mechanical models shown in Figures
324: \ref{fig:sprgms}(a) and (c), respectively.
325:
326: From the discussion above, one concludes that the polarization
327: energy and the associated losses depend intrinsically upon the
328: specific microscopic model giving rise to the susceptibility
329: $\chi$. However, these allocations cannot be relevant
330: macroscopically: for example, within the phenomenological
331: framework in which $\chi$ is introduced, the spatial and temporal
332: evolution of the various fields depend only upon this particular
333: piece of information, not upon its various representations. In
334: this circumstance in which the system dynamics is completely
335: determined by some piece of information (e.g. $\chi$), to say that
336: some other piece of information is important in order to establish
337: some particular notion of energy allocation is to admit that that
338: notion of energy allocation is irrelevant to those all-ready
339: specified dynamics. We establish a notion of energy allocation
340: that is determined uniquely by $\chi$ and, so, is relevant to
341: system dynamics. In particular it provides a precise notion of the
342: maximum possibility for the field to recover energy from the
343: medium in, for example, the borrow-return process mentioned at the
344: beginning of this article.
345:
346: In this article, we introduce the \emph{irreversible energy}
347: density $u_{irrev.}(t)$. At any given time $t$, it is defined to
348: be the infemum of all possible [LL] asymptotic losses
349: $u_{int.}(+\infty)$, this extrema being realized over all possible
350: future evolutions of the electric field $E$, holding its past
351: evolution fixed. Temporarily using notation emphasizing that the
352: internal energy at time $t$ depends not only on this time, but
353: also on the history of the instigating electric field $E$ up until
354: that time, one writes
355:
356: \begin{align}
357: u_{irrev.}[E](t)=u_{irrev.}[E_t^{-}](t)
358: &:=\underset{E_t^{+}}{\inf} u_{int.}[E_t^{-}+E_t^{+}](+\infty).
359: \label{Eq:uirrev}
360: \end{align}
361:
362: Here $E_t^{-}$ denotes the electric field time series $E(\tau)$
363: with its $t$-future ($\tau > t$) eliminated. Similarly, $E_t^{+}$
364: denotes an appended electric field time series $E(\tau)$ with its
365: $t$-past ($\tau < t$) eliminated. For a passive dielectric, this
366: infemum exists and, so, is unique \cite{AMEREM}. In particular,
367: and is shown later in Eq. (\ref{Eq:RI}), it does not depend upon
368: an explicit representation for $\chi$.
369:
370: Almost tautologically, from definition (\ref{Eq:uirrev}), it
371: follows that $u_{irrev.}(t)$ can never decrease as time increases.
372: Thus, at any given time $t$, it quantifies a component of the
373: medium internal energy $u_{int.}(t)$ that will, under all
374: circumstances, remain in and be dissipated by the medium. That is
375: it specifies a component of the medium internal energy that cannot
376: under any circumstances be returned to the field. Moreover, since
377: it is defined in terms of an infemum, i.e. a greatest lower bound,
378: all notions of loss greater than this value are too pessimistic:
379: at any given time $t$, there always exists a future medium-field
380: interaction creating less eventual energy loss to the medium than
381: any value greater than that specified by $u_{irrev.}(t)$. This is
382: true regardless of how small this overestimation is. Consequently,
383: within the phenomenological, macroscopic framework in which $\chi$
384: dictates the system dynamics, this quantity uniquely records the
385: energetic irreversibility generated within this dissipative
386: system.
387:
388: In Fig. \ref{fig:free} the dot-dashed curve specifies the
389: irreversible energy for the case considered, obviously valid for
390: either of the two $\chi$-equivalent physical systems. Note that in
391: the figure it is never exceeded by the losses in the sense of
392: Loudon, Barash and Ginzburg. Indeed one can prove that this
393: relationship must hold between the macroscopically relevant
394: irreversible energy and any notion of loss specified
395: microscopically. Equivalently one determines that energy (the
396: ability to do work) specified macroscopically cannot exceed any
397: such microscopic notion. Further, one concludes that the former is
398: almost always strictly less than the latter because of incoherence
399: among the system's microscopic, energy containing elements. The
400: former statement (the one regarding losses) can be obtained by
401: repeated application of the following theorem: (As in the
402: monotonicity of $u_{irrev.}(t)$, the theorem follows almost
403: tautologically from the definition.)
404:
405: \begin{align}
406: u_{irrev.}[E;\chi_1+\chi_2](t) \geq
407: u_{irrev.}[E;\chi_1](t)+u_{irrev.}[E;\chi_2](t),
408: \label{Eq:entropy}
409: \end{align}
410:
411: with strict inequality holding almost always for nontrivial
412: $\chi_1$ and $\chi_2$. Eq. (\ref{Eq:entropy}) demonstrates that in
413: "additive" processes (i.e. generating media mixtures),
414: irreversibility is generated. By repeated (additive) subdivisions
415: of a macroscopically relevant susceptibility $\chi$ into pieces
416: $\chi_i$, one may obtain microscopic representations of the medium
417: response. If the elements are "simple" enough (to be quantified
418: later) $u_{irrev.}[E;\chi_i]$ will be equivalent to the
419: parameter-dependent notion of loss suggested by Loudon, and
420: Ginzburg and Barash. Indeed the notion of loss and energy
421: specified by Loudon happen to agree with the
422: macroscopically/phenomenologically relevant notion herein
423: introduced in the cases he considered-- single Lorentz oscillator
424: media. [This is not the case for (competing) multiple oscillator
425: media considered by Barash and Ginzburg, as demonstrated by figure
426: \ref{fig:free}.]
427:
428: The difference between the current internal energy and the current
429: irreversible energy gives the \emph{reversible energy}
430: $u_{rev.}(t)$:
431:
432: \begin{align}
433: u_{rev.}(t):= u_{int.}(t)-u_{irrev.}(t).
434: \label{Eq:urev}
435: \end{align}
436:
437: The reversible energy gives the least upper bound on the amount of
438: energy that the medium can relinquish after time $t$: any amount
439: greater than this value, no matter how small the discrepancy,
440: overestimates the ability of the medium's microscopic,
441: energy-containing components, say, to organize themselves and do
442: useful, macroscopic work. Obviously the difference between the
443: (piece-wise constant) internal energy plotted in Fig.
444: \ref{fig:free}, and the dot-dashed curve in that figure, gives
445: this dynamical notion of the possibility for the medium to do work
446: (against the field) for the example considered.
447:
448: Using Eq. (\ref{Eq:uint}), one immediately shows that
449: $u_{int.}(t)$ is constant after the electric field ceases. From
450: definition (\ref{Eq:urev}), and the theorem regarding the
451: monotonicity of $u_{irrev.}(t)$, it follows that $u_{rev.}(t)$ can
452: never increase after such time, i.e. after the electric field
453: quits subsidizing its existence by doing work against the
454: polarized medium. We will show that $u_{rev.}(t)$ is never
455: negative. Consequently, one sees that when the system becomes
456: energetically closed, the reversible energy behaves like a
457: dynamical system free energy (density), equivalently like a system
458: Lyapunov function (density). For this reason, and for the
459: microscopic consideration mentioned, in particular the
460: entropy-generation-like statement embodied in Eq.
461: (\ref{Eq:entropy}), we will also designate the reversible energy
462: as the (medium-field) \emph{free energy}.
463:
464: We finish with a formula demonstrating how the macroscopic loss
465: (and so the free energy (via \ref{Eq:urev})), can be calculated.
466: (In particular how the irreversible energy plotted in figure
467: (\ref{fig:free}) can be generated.) This formula is obtained by
468: applying a variational principle to the definition
469: (\ref{Eq:uirrev}), and solving the resulting
470: \emph{Riemann-Hilbert} problem. One finds that, for passive,
471: causal dielectrics,
472:
473: \begin{eqnarray}
474: \begin{aligned}
475: u_{irrev.}[E](t) &=\frac{\lambda}{2 \pi}
476: \int\limits_{-\infty}^{t}\left|\int\limits_{-\infty}^{+\infty}
477: \frac{-i\omega\chi(\omega)E_{\tau}(\omega)}{\phi_+(\omega)}e^{-i\omega
478: \tau} \, d\omega \right|^2 \, d\tau \label{Eq:RI}
479: %\\
480: %\intertext{where}
481: \end{aligned}
482: \end{eqnarray}
483: where
484: \begin{eqnarray}
485: \begin{aligned}
486: \lambda &= \lim_{\omega \to \infty}
487: \frac{\textup{Im}[\chi(\omega)]}{\omega\left|\chi(\omega)\right|^2},\\
488: \phi_+(\omega) &= \lim_{\epsilon \to 0^+} \exp\left[\frac{-1}{2\pi
489: i}\int\limits_{-\infty}^{+\infty}\frac{\log\frac{\textup{Im}[\chi(\omega^{\prime})]}
490: {\lambda\omega^{\prime}\left|\chi(\omega^{\prime})\right|^2}}{\omega^{\prime}-\left(\omega+i\epsilon\right)}
491: \,d\omega^{\prime} \right]
492: \end{aligned}
493: \end{eqnarray}
494:
495: Here we introduce the \emph{instantaneous spectrum at time $t$},
496: $E_{t}(\omega)$. It is the Fourier transform of $E_t^{-}$ (see the
497: lines following definition (\ref{Eq:uirrev})). We also introduce
498: the \emph{medium complexity factor} $\phi_+(\omega)$. Its
499: deviation from unity gives a measure of the effective
500: macroscopic/phenomenological incoherence of possible microscopic,
501: energy containing elements. Media for which $\phi_+(\omega)$ is
502: identically unity we call \emph{simple} media, the rest we call
503: \emph{complex}. In the case that the susceptibility $\chi$
504: corresponds to a single Lorentz oscillator medium, as in the case
505: considered by Loudon, $\phi_+(\omega)$ is identically one, and Eq.
506: (\ref{Eq:RI}) reduces to
507:
508: \begin{eqnarray}
509: \begin{aligned}
510: u_{irrev.}[E](t) &=\lambda \int\limits_{-\infty}^{t} \dot
511: P^2(\tau) \, \, d\tau \label{Eq:RIL}
512: %\\
513: %\intertext{where}
514: \end{aligned}
515: \end{eqnarray}
516: where, then, $\lambda$ is determined in terms of the
517: phenomenological damping parameter $\gamma$ and the effective
518: plasma frequency $\omega_p$:
519: \begin{eqnarray}
520: \begin{aligned}
521: \lambda &= \gamma/\omega_p^2.
522: \end{aligned}
523: \end{eqnarray}
524:
525: In such case, then, and as claimed by Loudon, the relevant losses
526: correspond to the frictional losses generated by the single
527: Lorentz oscillator.
528:
529: \section{Summary}
530:
531: We have introduced notions of free energy and losses relevant to
532: the macroscopic behavior of passive, linear dielectric media.
533: These notions are relevant in the same regime in which the
534: macroscopic Maxwell equations are themselves relevant, i.e. in the
535: regime in which that theory specifies all measurable dynamics. In
536: particular, the macroscopic theory introduced is relevant to the
537: regimes in which the macroscopic, energy barrow-return process
538: describes the production of slow and fast light in passive,
539: linear, temporally dispersive media. Further communications will
540: describe the precise evolutions of the ideal medium-field
541: interactions giving rise to maximum energy recovery, i.e. those
542: "recovery" fields suggested by the variational definition
543: (\ref{Eq:uirrev}). The nature of the analogy of such recovery
544: fields in dissipative media with reversible processes in the
545: thermodynamic setting will be analyzed. Finally, the nonintuitive
546: evolution of the medium-field interaction on such recovery fields
547: in complex media will be exposed, e.g. the failure of monotonicity
548: in the evolution of the internal energy to its minimum value, even
549: on ideal recovery fields, and the identification of such as a
550: measure of the level of macroscopic disorder created by the
551: existence of many, competing degrees of freedom.
552:
553: \begin{acknowledgments}
554: Conversations with Kurt Oughstun, Joseph Eberly, Peter Milonni,
555: Aephraim Steinberg, Raymond Chiao, and Michael Fleischhauer are
556: gratefully acknowledged. This research was supported in part by
557: the National Science Foundation under Grant No. PHY99-07949.
558: \end{acknowledgments}
559:
560:
561:
562:
563:
564:
565:
566:
567:
568:
569: % If in two-column mode, this environment will change to single-column
570: % format so that long equations can be displayed. Use
571: % sparingly.
572: %\begin{widetext}
573: % put long equation here
574: %\end{widetext}
575:
576: % figures should be put into the text as floats.
577: % Use the graphics or graphicx packages (distributed with LaTeX2e)
578: % and the \includegraphics macro defined in those packages.
579: % See the LaTeX Graphics Companion by Michel Goosens, Sebastian Rahtz,
580: % and Frank Mittelbach for instance.
581: %
582: % Here is an example of the general form of a figure:
583: % Fill in the caption in the braces of the \caption{} command. Put the label
584: % that you will use with \ref{} command in the braces of the \label{} command.
585: % Use the figure* environment if the figure should span across the
586: % entire page. There is no need to do explicit centering.
587:
588: % \begin{figure}
589: % \includegraphics{}%
590: % \caption{\label{}}
591: % \end{figure}
592:
593: % Surround figure environment with turnpage environment for landscape
594: % figure
595: % \begin{turnpage}
596: % \begin{figure}
597: % \includegraphics{}%
598: % \caption{\label{}}
599: % \end{figure}
600: % \end{turnpage}
601:
602: % tables should appear as floats within the text
603: %
604: % Here is an example of the general form of a table:
605: % Fill in the caption in the braces of the \caption{} command. Put the label
606: % that you will use with \ref{} command in the braces of the \label{} command.
607: % Insert the column specifiers (l, r, c, d, etc.) in the empty braces of the
608: % \begin{tabular}{} command.
609: % The ruledtabular enviroment adds doubled rules to table and sets a
610: % reasonable default table settings.
611: % Use the table* environment to get a full-width table in two-column
612: % Add \usepackage{longtable} and the longtable (or longtable*}
613: % environment for nicely formatted long tables. Or use the [H]
614: % placement option to break a long table (with less control than
615: % in longtable).
616: % \begin{table}%[H] add [H] placement to break table across pages
617: % \caption{\label{}}
618: % \begin{ruledtabular}
619: % \begin{tabular}{}
620: % Lines of table here ending with \\
621: % \end{tabular}
622: % \end{ruledtabular}
623: % \end{table}
624:
625: % Surround table environment with turnpage environment for landscape
626: % table
627: % \begin{turnpage}
628: % \begin{table}
629: % \caption{\label{}}
630: % \begin{ruledtabular}
631: % \begin{tabular}{}
632: % \end{tabular}
633: % \end{ruledtabular}
634: % \end{table}
635: % \end{turnpage}
636:
637: % Specify following sections are appendices. Use \appendix* if there
638: % only one appendix.
639: %\appendix
640: %\section{}
641:
642: % If you have acknowledgments, this puts in the proper section head.
643: %\begin{acknowledgments}
644: % put your acknowledgments here.
645: %\end{acknowledgments}
646:
647: % Create the reference section using BibTeX:
648: \bibliography{PRL.bbl}
649:
650: \end{document}
651: %
652: % ****** End of file template.aps ******
653: