1:
2: \documentclass[twocolumn,aps,prd]{revtex4}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \usepackage{amsfonts}
5: \usepackage{amsmath}
6: \usepackage{amssymb}
7: \usepackage{graphicx}
8:
9:
10: \newcommand{\sech}{\mathrm{sech}}
11: \newcommand{\be}{\begin{equation}}
12: \newcommand{\ee}{\end{equation}}
13: \newcommand{\ben}{\begin{eqnarray}}
14: \newcommand{\een}{\end{eqnarray}}
15: \newcommand{\bb}{\bibitem}
16: \newcommand{\ov}{\overline}
17: \newcommand{\wt}{\widetilde}
18:
19: \begin{document}
20:
21: \title{Lorentz violating effects on a quantized two-level system}
22: \author{M. M. Ferreira Jr.}
23: \affiliation{Universidade Federal do Maranh\~ao (UFMA), Departamento de F\'\i sica \\
24: Campus Universit\'ario do Bacanga, 65085-580, S\~ao Lu\'\i s, Maranh\~ao,
25: Brazil, e-mail: manojr@ufma.br}
26: \author{A.R. Gomes}
27: \affiliation{Departamento de Ci\^encias Exatas, Centro Federal de Educa\c c\~ao
28: Tecnol\'ogica do Maranh\~ao, 65025-001 S\~ao Lu\'\i s, Maranh\~ao, Brazil,
29: e-mail: {argomes@pq.cnpq.br}}
30:
31: \begin{abstract}
32: In this work, we consider the effects of the Lorentz-violating (LV)\ term $%
33: v_{\mu }\overline{\psi }\gamma ^{\mu }\psi $ belonging to the fermion sector
34: of the extended standard model on the dynamics of a quantum two-level
35: system. We examine how its non-relativistic counterpart, ${(\mathbf{p}-e%
36: \mathbf{A})\cdot }\mathbf{v}/{m_{e},}$ affects the Rabi oscillations of a
37: two-level atom coupled with a quantum cavity electromagnetic field. Taking
38: an initial coherent field state in a resonant cavity, it was found that the
39: LV\ background increases the Rabi frequency and the time interval between
40: collapses and revivals of the population inversion function. It was found
41: that initial field states with low mean number of photons are better probes
42: in order to establish more stringent upper bounds on the background
43: magnitude. In particular, for an initial vacuum state in the cavity the
44: upper limit $\text{v}_{x}<10^{-10}eV$ was attained.
45: \end{abstract}
46:
47: \maketitle
48:
49: %\pacs{xxx,yyy,zzz.}
50:
51: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
52:
53: \section{introduction}
54:
55: In the latest years, Lorentz violation (LV)\ in physical systems has been
56: investigated in connection with a possible breakdown of this symmetry at the
57: Planck scale. Since the demonstration that spontaneous breaking of Lorentz
58: symmetry may occur in the context of string theory \cite{Samuel}, small
59: violations of Lorentz covariance in low-energy systems have been searched as
60: a remanent effect of LV at the Planck scale. Naturally, this is a relevant
61: issue, once the Planck scale physics is entirely unknown yet. Nowadays, most
62: LV\ investigations have been conducted into the framework of the Standard
63: Model Extension (SME) \cite{Colladay}, wherein LV is incorporated in all
64: sectors of interaction and governed by tensor coefficients generated as
65: vacuum expectation values of tensor quantities of the original symmetric
66: theory. In such a model, Lorentz breaking takes place only in the particle
67: frame, where such coefficients behave as a set of numbers. In the observer
68: frame, the Lorentz covariance remains valid.
69:
70: In the framework of the SME, LV has been intensively investigated concerning
71: mainly the photon \cite{Photon} and fermion sectors with many different
72: purposes, involving radiative corrections \cite{Radiative}, topological
73: effects \cite{Topological}, CPT probing experiments \cite{CPT}, hydrogen
74: spectrum \cite{Hydrogen}, and general aspects \cite{general}. Atomic and
75: optical systems \cite{Optical} (including resonant cavities) have been used
76: as a laboratory to test the limits of Lorentz covariance, implying stringent
77: bounds on the LV coefficients.
78:
79: The fermion Lagrangian of the SME includes two Lorentz and CPT-odd terms, $%
80: v_{\mu }\overline{\psi }\gamma ^{\mu }\psi ,b_{\mu }\overline{\psi }\gamma
81: _{5}\gamma ^{\mu }\psi ,$ with $v_{\mu },b_{\mu }$ being the LV coefficients
82: generated as vacuum expectation values of tensor quantities belonging to the
83: underlying high energy theory. In a very recent work, it was analyzed the
84: effect of these two terms on a semi-classical two-level system \cite{Semi},
85: being observed that they yield alterations on its dynamics, inducing
86: sensitive modifications on the usual inversion population function and
87: quantum transitions even in the absence of an external electromagnetic field.
88:
89: In the present work we consider the effect of the LV term $v_{\mu }\bar{\psi}%
90: \gamma ^{\mu }\psi $ on a quantum atomic system. The starting point is the
91: extended Lagrangian
92: \begin{equation}
93: \mathcal{L}%
94: %TCIMACRO{\U{b4}}%
95: %BeginExpansion
96: {\acute{}}%
97: %EndExpansion
98: \mathcal{=L}_{Dirac}-v_{\mu }\overline{\psi }\gamma ^{\mu }\psi \label{L1}
99: \end{equation}%
100: where $\mathcal{L}_{Dirac}$ is the usual Dirac Lagrangian ($\mathcal{L}%
101: _{Dirac}=\frac{1}{2}i\overline{\psi }\gamma ^{\mu }\overleftrightarrow{%
102: D }_{\mu }\psi -m_{e}\overline{\psi }\psi )$. It yields the following
103: nonrelativistic Hamiltonian
104: \begin{equation}
105: H=H_{Pauli}+\biggl[-\frac{(\mathbf{p}-e\mathbf{A})\cdot \mathbf{v}}{m_{e}}%
106: \biggr].
107: \end{equation}%
108: The purpose here is to examine the effects implied by the Lorentz-violating
109: Hamiltonian in the Rabi nutation of a single atom in the vacuum and in a
110: weak coherent field established in a resonant cavity. Using the amplitude
111: coefficient method, the Schr\"{o}dinger equation is taken as starting point
112: to obtain differential equations for the amplitude coefficients. These
113: equations govern the dynamics of the two-level system and allow to read the
114: effects induced by the LV background on it. After some approximation, it is
115: derived a system of two coupled typical harmonic oscillator differential
116: equations, whose solution leads to a modified expression for the inversion
117: population function. It then reveals that the Rabi frequency, the collapse $%
118: \left( t_{c}\right) $ and revival $\left( t_{r}\right) $ times all increase
119: with the background magnitude. At the same time, the revival packages become
120: larger and more distant from each other. Considering that quantum
121: experiments present a sensitivity of 1 part in $10^{10}$, an upper bound $%
122: \left( \text{v}_{x}\leq 10^{-10}eV\right) $ for the background is
123: established.
124:
125: This paper is outlined as follows. In Sec.\textbf{\ }II, it is presented a
126: brief resume on the interaction of a two-level system with a quantized
127: monochromatic field. In Sec. III, the Lorentz violating effects stemming
128: from the coupling $v_{\mu }\overline{\psi }\gamma ^{\mu }\psi $ on the
129: two-level system are properly examined by means of the amplitude coefficient
130: method. Special attention is paid to the LV effects on the Rabi frequency,
131: collapse and revival times. In Sec. IV, we present our Conclusion and final
132: remarks.
133:
134: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
135:
136: \section{Interaction of a two-level atom with a single mode field}
137:
138: \label{sect2}
139:
140: First of all, we review a standard result concerning the interaction
141: between a two-level atom with a quantized field (in the absence of the
142: background $\mathbf{v)}$. The two levels are identified as $|a\rangle $ and $%
143: |b\rangle $ with energy $(1/2)\hbar \omega $ and $-(1/2)\hbar \omega ,$
144: respectively. This system is described by the Hamiltonian in the Schr\"{o}%
145: dinger representation (See Ref. \cite{scully}, Chap. 6):
146: \begin{equation}
147: \hat{H}=\hat{H}_{atom}+\hat{H}_{field}+\hat{H_{1}}, \label{H1}
148: \end{equation}%
149: where $\hat{H}_{atom}=\hbar \omega _{a}|a\rangle +\hbar \omega _{b}|b\rangle
150: $ is the atomic Hamiltonian, \ $\hat{H}_{field}=$ $\sum_{k}\hbar \nu _{k}(%
151: \hat{a}^{\dagger }\hat{a}+1/2)$ is the radiation field Hamiltonian, $\hat{%
152: H_{1}}=-e\mathbf{r}\cdot \mathbf{E}$ describes the atom-field interaction in
153: the dipole approximation, and $\hat{a},\hat{a}^{\dagger }$ are the photon
154: destruction and creation operators. Here, $\mathbf{r}$ is the position
155: vector of the electron and $\mathbf{E}$ is the radiation field that
156: interacts with the atom,
157: \ben
158: \mathbf{E=}\sum\limits_{k}E_{0k}(\hat{a}e^{-i\nu
159: _{k}t}+\hat{a}^{\dagger }e^{i\nu _{k}t})\hat{\epsilon}_{k},
160: \een
161: with $\hat{%
162: \epsilon}_{k}$ being the polarization vector and $E_{0k}$ being the
163: amplitude of the mode of frequency $\nu _{k}.$ Such amplitude is given by $%
164: E_{0k}=\sqrt{\hbar \nu _{k}/(2\epsilon _{0}\mathcal{V})},$ where $\mathcal{V}
165: $ is the cavity effective volume. This normalization factor is obtained by
166: equating the Fock states energy with the integral over space of the
167: expectation value of the electromagnetic energy density. In this work, we
168: will consider the interaction of a single-mode field of frequency $\nu $
169: with the two-level atom, so from now on we write $\hat{H}_{field}=\hbar \nu (%
170: \hat{a}^{\dagger }\hat{a}+1/2)$ and $\mathbf{E=}E_{0}(\hat{a}e^{-i\nu t}+%
171: \hat{a}^{\dagger }e^{i\nu t})\hat{\epsilon}_{k}.$
172:
173: Now, the Hamiltonian (\ref{H1}) can be read as $\hat{H}=\hat{H}_{0}+\hat{%
174: H_{1}},$ where $\hat{H}_{0}=\hat{H}_{atom}+\hat{H}_{field}$ \ plays the role
175: of the unperturbed interaction and $\hat{H_{1}}=-e\mathbf{r}\cdot \mathbf{E}$
176: can be viewed as a small perturbation. The approach adopted here is
177: developed in the interaction picture, wherein the state vectors evolve with $%
178: \hat{H_{1}}$ whereas the operators evolve with $\hat{H_{0}}$ \cite{foot1}.
179: In such a picture, the interaction operator $H_{1}$ is to be written as $%
180: \hat{H_{1}}_{I}(t)=e^{iH_{0S}t/\hbar }\hat{H_{1}}_{S}e^{-iH_{0S}t/\hbar },$
181: where $\hat{H_{1}}_{S}$ stands for the atom-field interaction in the Schr%
182: \"{o}dinger picture, in which the operators do not present time dependence
183: (see also Ref. \cite{loudon}, p. 187). For this reason, the time dependence
184: of $\hat{\text{ }H_{1}}$ will be dropped out from now on.
185:
186: In order to evaluate $\hat{H_{1}}$ in a more suitable form, we use the atom
187: transition operators, $\hat{\sigma}_{ij}=|i\rangle \langle j|,$ where $%
188: |i\rangle $ represents a complete set of energy eigenstates, so that $1=\sum
189: |i\rangle \langle i|$. In our two-level case, $|i\rangle =|a\rangle $ or $%
190: |b\rangle ,$ obviously. Considering it, we obtain:
191: \begin{equation}
192: \hat{H_{1}}=\sum_{ij}g^{ij}\hat{\sigma}_{ij}(\hat{a}+\hat{a}^{\dagger }),
193: \end{equation}%
194: where the electric field was evaluated at $t=0$ (due to the choice of the
195: Schr\"{o}dinger representation), $g^{ij}=-e(\mathbf{P}_{ij}\cdot \hat{%
196: \epsilon}_{k})E_{0k}/\hslash ,$ and $e\mathbf{P}_{ij}=e\langle i|\mathbf{r}%
197: |j\rangle $ is the transition matrix element of the electric dipole moment.
198: Supposing that $\mathbf{P}_{ab}=\mathbf{P}_{ba},g^{ab}=g^{ba}=g,$ the
199: interaction $\hat{H_{1}}$ takes the form $\hat{H_{1}}=g(\hat{\sigma}_{ab}+%
200: \hat{\sigma}_{ba})(\hat{a}+\hat{a}^{\dagger }).$ The terms $\hat{\sigma}_{ab}%
201: \hat{a}^{\dagger }$ and $\hat{\sigma}_{ba}\hat{a}\ $should be neglected.
202: Indeed, the term $\hat{\sigma}_{ab}\hat{a}^{\dagger }$ induces an atomic
203: transition from the ground state ($|b\rangle $) to the excited state ($%
204: |a\rangle )$ while a photon of frequency $\nu $ is emitted. The term $\hat{%
205: \sigma}_{ba}\hat{a}$ \ implies an atomic transition from the excited state ($%
206: |a\rangle $) to the ground state ($|b\rangle $) while a photon of frequency $%
207: \nu $ is absorbed. Both processes do not conserve energy. The exclusion of
208: the non-conserving energy terms is equivalent to the rotating wave
209: approximation (RWA). In the semiclassical theory it takes place a similar
210: fact: the non-resonant terms are neglected. We now introduce the notation: $%
211: \hat{\sigma}_{_{+}}=\hat{\sigma}_{ab}=|a\rangle \langle b|$, $\hat{\sigma}%
212: _{_{-}}=\hat{\sigma}_{ba}=|b\rangle \langle a|,$ so that the
213: energy-conserving Hamiltonian takes the form $\hat{H}=\hat{H}_{0}+\hat{H}_{1}
214: $, with
215: \begin{eqnarray}
216: \hat{H_{0}} &=&\hbar \nu \hat{a}^{\dagger }\hat{a}+\frac{1}{2}\hbar \omega
217: \hat{\sigma}_{z}, \label{H_single1} \\
218: \hat{H_{1}} &=&\hbar g(\hat{\sigma}_{_{+}}\hat{a}+\hat{\sigma}_{_{-}}\hat{a}%
219: ^{\dagger }). \label{H_single2}
220: \end{eqnarray}%
221: The operator $\hat{\sigma}_{_{+}}$ leads the atom from state $|b\rangle $ to
222: state $|a\rangle $, whereas $\hat{\sigma}_{_{-}}$ makes the inverse
223: operation. We should also define
224: \begin{eqnarray}
225: \hat{\sigma}_{z} &=&|a\rangle \langle a|-|b\rangle \langle b|,\hat{\sigma}%
226: _{x}=(\hat{\sigma}_{_{+}}+\hat{\sigma}_{_{-}}), \label{spin} \\
227: \hat{\sigma}_{_{y}} &=&-i(\hat{\sigma}_{_{+}}-\hat{\sigma}_{_{-}}), \label{spin_y}
228: \end{eqnarray}%
229: operators which fulfill the Pauli algebra ($\left[ \hat{\sigma}_{i},\hat{%
230: \sigma}_{j}\right] =2i\epsilon _{ijk}\hat{\sigma}_{k}).$
231:
232: Considering the case the electric field is linearly polarized in the
233: x-direction and $\mathbf{P}_{ab}$ is real \cite{foot2}, we can write
234: \begin{equation}
235: g=-\frac{e{P}_{ab}E_{0}}{\hbar }. \label{g}
236: \end{equation}%
237: The normalization factor is equal to
238: \begin{equation}
239: E_{0}=\sqrt{\hbar \nu /(2\epsilon _{0}\mathcal{V)}}. \label{eqE0}
240: \end{equation}%
241: Experimentally the cavity set-up provides a precise description for the
242: atomic dynamics even with the atom-field Hamiltonian $\hat{H_{1}}$ given by
243: Eq. (\ref{H_single2}), since the interaction with a single mode dominates
244: the evolution.
245:
246: In the interaction picture, the interaction potential $\hat{V}$ is defined
247: as
248: \begin{equation}
249: \hat{V}=\hat{U}_{0}^{\dagger }(t)\hat{H}_{1S}\hat{U}_{0}(t), \label{V}
250: \end{equation}%
251: with $\hat{H}_{1S}$ being the Schr\"{o}dinger (time independent)
252: representation of $\hat{H}_{1S}$, and $\hat{U}_{0}(t)=e^{-i\hat{H_{0}}%
253: t/\hbar }$ is the time evolution operator in this picture. In order to
254: evaluate $\hat{V},$ the Baker-Campbell-Hausdorff lemma is applied and
255: yields:
256: \begin{eqnarray}
257: e^{i\nu \hat{a}^{\dagger }\hat{a}t}\hat{a}e^{-i\nu \hat{a}^{\dagger }\hat{a}%
258: t} &=&\hat{a}e^{-i\nu t}, \label{BCHa} \\
259: e^{i\nu \hat{a}^{\dagger }\hat{a}t}\hat{a}^{\dagger }e^{-i\nu \hat{a}%
260: ^{\dagger }\hat{a}t} &=&\hat{a}^{\dagger }e^{i\nu t}, \label{BCHadag} \\
261: e^{\frac{1}{2}i\omega \hat{\sigma}_{z}t}\hat{\sigma}_{\pm }e^{-\frac{1}{2}%
262: i\omega \hat{\sigma}_{z}t} &=&\hat{\sigma}_{_{\pm }}e^{\pm i\omega t}.
263: \label{b}
264: \end{eqnarray}%
265: Replacing the former relations and Eq. (\ref{H_single2}) on Eq. (\ref{V}),
266: we get
267: \begin{equation}
268: \hat{V}=\hbar g(\hat{\sigma}_{_{+}}\hat{a}e^{i\Delta t}+\hat{a}^{\dagger }%
269: \hat{\sigma}_{_{-}}e^{-i\Delta t}),
270: \end{equation}%
271: with $\Delta =\left( \omega -\nu \right) $. In this representation, the
272: wavefunction $|\psi _{I}(t)\rangle =\widehat{U}_{0}^{\dagger }(t)|\psi
273: _{S}(t)\rangle $ is a linear combination of the two atomic states with
274: arbitrary number of photons $(n)$ in the cavity ($|a,n\rangle $ and $%
275: |b,n\rangle )$. So we have
276: \begin{equation}
277: |\psi _{I}(t)\rangle =\sum_{n=0}^{\infty }[c_{a,n}|a,n\rangle
278: +c_{b,n}|b,n\rangle ].
279: \end{equation}%
280: The Schr\"{o}dinger equation in the interaction picture $i\hbar |\dot{\psi}%
281: _{I}(t)\rangle =\hat{V}|\psi _{I}(t)\rangle $ leads to the following system
282: of coupled differential equations:
283: \begin{eqnarray}
284: \dot{c}_{a,n} &=&-ig\sqrt{n+1}e^{i\Delta t}c_{b,n+1}, \\
285: \dot{c}_{b,n+1} &=&-ig\sqrt{n+1}e^{-i\Delta t}c_{a,n},
286: \end{eqnarray}%
287: which can be easily solved. Considering the atom initially in the excited
288: state $|a\rangle $ and the field with a distribution $c_{n}(0)$ of photons,
289: we can write $c_{a,n}(0)=c_{n}(0)$ and $c_{b,n+1}(0)=0$ to get (see Ref.
290: \cite{scully}, chap. 6):
291: \begin{eqnarray}
292: c_{a,n}(t) &=&c_{n}(0)e^{i\Delta t/2}[\cos (\gamma t)-\frac{i\Delta }{\Omega
293: _{n}}\sin (\gamma t)], \\
294: c_{b,n+1}(t) &=&-c_{n}(0)\frac{2ig\sqrt{n+1}}{\Omega _{n}}\sin (\gamma
295: t)e^{-i\Delta t/2},
296: \end{eqnarray}%
297: where $\Omega _{n}^{2}=\Delta ^{2}+4g^{2}(n+1)$ and $\gamma =\Omega _{n}/2$.
298: The atomic inversion function can be now defined as
299: \begin{equation}
300: W(t)=\sum_{n=0}^{\infty }(|c_{a,n}(t)|^{2}-|c_{b,n}(t)|^{2}),
301: \end{equation}%
302: Here, the discrete character of the sum over the number of photons is crucial
303: for the observation of periodic collapses and revivals of the inversion
304: function as a pure quantum effect (see Ref.\cite{walls}, chap. 10). For the
305: particular case of resonance ($\Delta =0$) we have
306: \begin{equation}
307: W(t)=\sum_{n=0}^{\infty }\rho _{nn}(0)\cos (2\sqrt{n+1}gt), \label{W_coh}
308: \end{equation}%
309: where $\rho _{nn}(0)=|c_{n}(0)|^{2}$ is the probability that there are $n$
310: photons in the cavity at $t=0$.
311:
312: Considering an initial coherent state in a cavity with medium number of
313: photons $\bar{n}$ we obtain
314: \begin{equation}
315: W(t)=\sum_{n=0}^{\infty }\frac{\bar{n}^{n}e^{-\bar{n}}}{n!}\cos \text{(}2%
316: \sqrt{n+1}gt\text{)}.
317: \end{equation}%
318: For $\bar{n}>>1$ (but not so large, in order to fulfill the RWA
319: condition $g\sqrt{\bar{n}}\ll \omega $ \cite{rwa}), with small variance $%
320: \Delta n,$ we recover the known Rabi nutation with frequency $\Omega _{\bar{n%
321: }}^{0}\simeq 2g\sqrt{\bar{n}}$. This is the semiclassical expression as
322: expected from the correspondence principle. For intermediate values of $\bar{%
323: n}$, this expression leads to the characteristic behavior of collapses and
324: revivals of the population inversion. This known effect appears mainly due
325: to the interference of the several oscillating patterns associated to
326: different photon numbers.
327:
328: In particular, at resonance and taking the vacuum ($\rho _{nn}(0)=\delta
329: _{n,0}$) as the initial state in the cavity, we find $W(t)=\cos (2gt)$. So,
330: Rabi oscillations with frequency $\Omega =2g$ occur due to spontaneous
331: emission, a typical quantum effect that does not occur in the semi-classical
332: system.
333:
334: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
335:
336: \section{Lorentz-violation effects}
337:
338: Now we will explore how these quantum fundamental effects can be affected
339: depending on the strength of the Lorentz-breaking coupling in Lagrangian (%
340: \ref{L1}). The Hamiltonian can now be written as
341: \begin{equation}
342: \hat{H}=\hat{H_{0}}+\hat{H_{1}}+\hat{H_{1}^{\prime }}+\hat{H_{2}^{\prime }},
343: \end{equation}%
344: where now we have the additional LV contributions: $\hat{H_{1}^{\prime }}=e%
345: \mathbf{A}\cdot \mathbf{v}/m_{e}$ and $\hat{H_{2}^{\prime }}=-\mathbf{p}%
346: \cdot \mathbf{v}/m_{e}$.
347:
348: First of all, we analyze the $\hat{H_{1}^{\prime }}=e\widehat{A}_{x}$v$%
349: _{x}/m_{e}$ contribution. After quantization, in the time-independent Schr%
350: \"{o}dinger representation, we can write
351: \begin{equation}
352: \hat{H_{1}^{\prime }}=\frac{ieE_{0}\text{v}_{x}}{m_{e}\nu }(-\hat{a}+\hat{a}%
353: ^{\dagger }). \label{eqH1pr}
354: \end{equation}%
355: In order to obtain the interaction potential $\hat{V}_{1}^{\prime }=\hat{U}%
356: _{0}^{\dagger }(t)\hat{H}_{1}^{\prime }\hat{U}_{0}(t)$, we use Eqs. (\ref%
357: {eqH1pr}), (\ref{BCHa}) and (\ref{BCHadag}), leading to
358: \begin{equation}
359: \hat{V}_{1}^{\prime }=\frac{ieE_{0}\text{v}_{x}}{m_{e}\nu }(-\hat{a}e^{-i\nu
360: t}+\hat{a}^{\dagger }e^{i\nu t}).
361: \end{equation}
362:
363: Now we turn to the contribution of $\hat{H_{2}^{\prime }}=-\widehat{p}_{x}$v$%
364: _{x}/m_{e}=-\widehat{\dot{x}}$v$_{x}$. The $\widehat{\dot{x}}$ operator can
365: be rewritten using the Heisenberg equation in the interaction picture as ${%
366: \widehat{H}}_{2}^{\prime }=-$v$_{x}/(i\hbar )[\widehat{x},\widehat{H}_{0}]$.
367: We found better to represent the $\hat{x}$ operator as
368: \begin{equation}
369: \hat{x}=-iP_{ab}\hat{\sigma}_{y}\hat{\sigma}_{z}, \label{x_op}
370: \end{equation}%
371: with $P_{ab}\equiv \langle a|\hat{x}|b\rangle =P_{ab}^{\ast }=P_{ba}$, where
372: the operators $\hat{\sigma}_{z},$ $\hat{\sigma}_{y}$ are defined in Eqs. (\ref%
373: {spin})-(\ref{spin_y}). If we represent the energy eigenstates in a vector form
374: \begin{equation}
375: |a\rangle =\left(
376: \begin{array}{c}
377: 1 \\
378: 0%
379: \end{array}%
380: \right) ,\,\,\,|b\rangle =\left(
381: \begin{array}{c}
382: 0 \\
383: 1%
384: \end{array}%
385: \right) ,
386: \end{equation}%
387: we can identify the operators $\hat{\sigma}_{x},\hat{\sigma}_{y},\hat{\sigma}%
388: _{z}$ with the Pauli matrices
389: \begin{equation}
390: \hat{\sigma}_{x}=\left(
391: \begin{array}{cc}
392: 0 & 1 \\
393: 1 & 0%
394: \end{array}%
395: \right) ,\,\,\,\hat{\sigma}_{y}=\left(
396: \begin{array}{cc}
397: 0 & -i \\
398: i & 0%
399: \end{array}%
400: \right) ,\,\,\,\hat{\sigma}_{z}=\left(
401: \begin{array}{cc}
402: 1 & 0 \\
403: 0 & -1%
404: \end{array}%
405: \right) .
406: \end{equation}%
407: Some remarks about the mathematical notation are worthy. Pauli matrix $%
408: \sigma _{z}$ was already used in Eq. (\ref{H_single1}) for the free
409: Hamiltonian as an economic way to describe the atom energy content. Hence,
410: the use of Pauli matrices in this way has no connection with spin operators
411: or spin states. These so-called pseudo-spin operators (Ref. \cite{walls}, p.
412: 203) are very useful to simplify the calculations and should not lead us to
413: misunderstandings concerning spin magnetic features of the atom states.
414:
415: In order to achieve a description in the interaction picture, we must apply
416: again the Baker-Campbell-Hausdorff lemma to obtain
417: \begin{equation}
418: e^{\frac{1}{2}i\omega \hat{\sigma}_{z}t}\hat{x}e^{-\frac{1}{2}i\omega \hat{%
419: \sigma}_{z}t}=-iP_{ab}[\sin (\omega t)\hat{\sigma}_{x}+\cos (\omega t)\hat{%
420: \sigma}_{y}]\hat{\sigma}_{z},
421: \end{equation}%
422: where Eq. (\ref{x_op}) was used. In this representation, the
423: interaction potential
424: \begin{equation}
425: \hat{V}_{2}^{\prime }=-\frac{\text{v}_{x}}{i\hbar }[e^{i\hat{H}_{0}t/\hbar }%
426: \hat{x}e^{-i\hat{H}_{0}t/\hbar }\hat{H}_{0}-\hat{H}_{0}e^{i\hat{H}%
427: _{0}t/\hbar }\hat{x}e^{-i\hat{H}_{0}t/\hbar }],
428: \end{equation}%
429: takes the form:
430: \begin{equation}
431: \hat{V}_{2}^{\prime }=i\text{v}_{x}P_{ab}\omega \lbrack \cos (\omega t)\hat{%
432: \sigma}_{x}-\sin (\omega t)\hat{\sigma}_{y}]\hat{\sigma}_{z}.
433: \end{equation}
434:
435: The Schr\"{o}dinger equation in the interaction picture, $i\hbar |\dot{\psi}%
436: _{I}(t)\rangle =(\hat{V}+\hat{V}_{1}^{\prime }+\hat{V}_{2}^{\prime })|\psi
437: _{I}(t)\rangle $, yields a system of coupled differential equations for the
438: probability amplitudes:
439: \begin{eqnarray}
440: \dot{c}_{a,n} &=&-ig\sqrt{n+1}e^{i\Delta t}c_{b,n+1} \notag \label{eom1} \\
441: &&+\frac{eE_{0}\text{v}_{x}}{m_{e}\hbar \nu }(-c_{a,n+1}\sqrt{n+1}e^{-i\nu
442: t}+c_{a,n-1}\sqrt{n}e^{i\nu t}) \notag \\
443: &&-\frac{\text{v}_{x}P_{ab}}{\hbar }\omega e^{i\omega t}c_{b,n}, \\
444: \dot{c}_{b,n+1} &=&-ig\sqrt{n+1}e^{-i\Delta t}c_{a,n} \notag \label{eom2}
445: \\
446: &&+\frac{eE_{0}\text{v}_{x}}{m_{e}\hbar \nu }(-c_{b,n+2}\sqrt{n+2}e^{-i\nu
447: t}+c_{b,n}\sqrt{n+1}e^{i\nu t}) \notag \\
448: &&+\frac{\text{v}_{x}P_{ab}}{\hbar }\omega e^{-i\omega t}c_{a,n+1}.
449: \end{eqnarray}%
450: with $n=0,1,...\infty $. In general, this system of infinite equations is of
451: difficult solution. Here, we will study the system at resonance ($\Delta =0$%
452: ).
453:
454: For large $n$ we have
455: \begin{eqnarray}
456: \dot{c}_{a,n} &=&-ig\sqrt{n}c_{b,n} \notag \\
457: &&+\frac{eE_{0}\text{v}_{x}}{m_{e}\hbar \nu }c_{a,n}\sqrt{n}2i\sin (\nu t)
458: \notag \\
459: &&-\frac{\text{v}_{x}P_{ab}}{\hbar }\omega e^{i\omega t}c_{b,n}, \\
460: \dot{c}_{b,n} &=&-ig\sqrt{n}c_{a,n} \notag \\
461: &&+\frac{eE_{0}\text{v}_{x}}{m_{e}\hbar \nu }c_{b,n}\sqrt{n}2i\sin (\nu t)
462: \notag \\
463: &&+\frac{\text{v}_{x}P_{ab}}{\hbar }\omega e^{-i\omega t}c_{a,n}.
464: \end{eqnarray}%
465: This must be compared with the differential equations for the coefficients
466: obtained in the semiclassical theory\cite{Semi}:
467: \begin{align}
468: \overset{\cdot }{a}(t)& =i(\Omega _{R}/2)b(t)+i\alpha _{0}a(t)\sin \nu
469: t-\beta _{0}\omega b(t)e^{i\omega t}, \label{a2} \\
470: \overset{\cdot }{b}(t)& =i(\Omega _{R}/2)a(t)+i\alpha _{0}b(t)\sin \nu
471: t+\beta _{0}\omega a(t)e^{-i\omega t}, \label{b2}
472: \end{align}%
473: where $\alpha _{0}=eE_{0}^{sc}$v$_{x}/(m_{e}\hbar \nu )$, $\beta _{0}=($v$%
474: _{x}P_{ab}/hbar)$, $\Omega_R$ is the Rabi frequency and $E_{0}^{sc}=\sqrt{2n\hbar \nu /\epsilon _{0}\mathcal{V}%
475: }$ is the semiclassical expression corresponding to the normalization factor
476: $E_{0}$. Note that $2\sqrt{n}E_{0}=E_{0}^{sc}$ and the correspondence
477: principle is verified.
478:
479: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
480:
481: We consider circular Rydberg atoms in a high Q cavity. In such atoms the
482: long radiative lifetime makes atomic relaxation negligible during the atom's
483: transit time across the cavity \cite{prl96}. Also a high Q cavity turns the
484: photon lifetime longer than the atom-cavity interaction time. Brune et al.
485: \cite{prl94} investigated the resonant effects of the vacuum in a cavity
486: mode. There the authors considered the transition frequency $\omega =51GHz$
487: for $n=50$ circular Rydberg atoms. The matrix element between the circular
488: levels 50 and 51 of a linear projection of the electric dipole on the
489: orbit's plane has the large value $eP_{ab}=1250ea_{0}$, where $a_{0}$ is the
490: Bohr radius. This means a very large classical radius and a very large
491: radiative decay time. The effective cavity volume was $0.7cm^{3}$. With it,
492: Eq. (\ref{eqE0}) provides a vacuum field amplitude at antinodes of $%
493: E_{0}=6.95\times 10^{-4}V/m$. These parameters correspond to the following
494: vacuum Rabi oscillation frequency $\Omega _{vac}=2g={2eP_{ab}E_{0}}/\hbar
495: =132kHz$, which implies $g=66kHz$ for the coupling constant. In these
496: estimates we are neglecting the usual spatial variation of the
497: electromagnetic field in the cavity. We will also consider that the applied
498: field frequency is near resonance $\left( \nu =\omega =51GHz\right) $.
499:
500: Now, we can use the above values for the parameters to estimate the relative
501: importance of the Lorentz-violating terms in Eqs. (\ref{eom1})-(\ref{eom2}).
502: In this way, we have $eE_{0}/(m_{e}\hbar \omega )=2\times
503: 10^{31}kg^{-1}m^{-1}$ and $\ P_{ab}\omega /{\hbar }=3\times
504: 10^{37}kg^{-1}m^{-1}$. Then, we note that for $\bar{n}\ll 10^{12}$ the
505: magnitude of the second term (corresponding to the influence of $-\mathbf{p}%
506: \cdot \mathbf{v}/m_{e}$) can be taken as much larger than the magnitude of
507: the first one (corresponding to $e\mathbf{A}\cdot \mathbf{v}/m_{e}$). At
508: resonance, both terms oscillate with the same frequency, but the smaller
509: amplitude argument is an enough reason to neglect the term stemming from $e%
510: \mathbf{A}\cdot \mathbf{v}/m_{e}$. The fact that such term does not imply
511: quantum effects at a first approximation is in agreement with the
512: semi-classical behavior \cite{Semi} associated with it. Indeed, it amounts
513: only at phase effects without altering the semi-classical population
514: inversion function. Neglecting this term, it is ascribed an explicit
515: gauge-independent character for the LV modifications.
516:
517: From these considerations the equations for the probability amplitudes can
518: be properly approximated as
519:
520: \begin{eqnarray}
521: \dot{c}_{a,n} &\simeq &-ig\sqrt{n+1}c_{b,n+1} \notag \label{eom3} \\
522: &&-\frac{\text{v}_{x}P_{ab}}{\hbar }\omega e^{i\omega t}c_{b,n}, \\
523: \dot{c}_{b,n+1} &\simeq &-ig\sqrt{n+1}c_{a,n} \notag \label{eom4} \\
524: &&+\frac{\text{v}_{x}P_{ab}}{\hbar }\omega e^{-i\omega t}c_{a,n+1}.
525: \end{eqnarray}
526:
527: Here, an interesting issue is to know if the high frequency of the
528: Lorentz-violating term can in fact provide an effective correction. In fact,
529: we remind that a high-frequency term proportional to $2\omega $ was
530: discarded due to the rotating-wave approximation. The exclusion of the
531: counter-rotating terms from the equations of motion is justified when the
532: frequency of the external modes most strongly interacting with the system is
533: very large compared to the strength $g$ of the interaction \cite{walls} (in
534: the present case $\omega =51GHz$ whereas $g=66kHz$). We have already seen
535: that such exclusion is equivalent to neglecting the non-energy-conserving
536: processes such as the excitation of an atom along with the emission of a
537: photon \cite{walls}. From this point of view the Lorentz-violating term
538: induced by the background has an energy non-conserving character, once the
539: Hamiltonian of Eq. (\ref{eqH1pr}) amounts at creation and annihilation of a
540: photon without a corresponding change on the atomic level. One possibility
541: is to consider physical situations where such an energy-violating terms
542: grows in importance (see \cite{mil94}, p. 151) and \cite{cook}. However, as
543: the terms depending on the background oscillate with frequency $\omega $
544: whereas the discarded terms from the RWA oscillate with $2\omega $, the
545: investigation of the influence of the background on the modified equations
546: of motion (Eqs. (\ref{eom3})-(\ref{eom4})) seems to be a sensible option.
547:
548: We can decouple those equations using the approximation that during the time
549: interval of measurement of Rabi nutation the oscillating terms are averaged
550: to zero [$\left\langle \cos (\omega t)\right\rangle =\left\langle \sin
551: (\omega t)\right\rangle =0]$. This gives
552: \begin{eqnarray}
553: \ddot{c}_{a,n} &\simeq &\biggl(-g^{2}{(n+1)}-\frac{\text{v}%
554: _{x}^{2}P_{ab}^{2}\omega ^{2}}{\hbar ^{2}}\biggr)c_{a,n}\,, \\
555: \ddot{c}_{b,n+1} &\simeq &\biggl(-g^{2}{(n+1)}-\frac{\text{v}%
556: _{x}^{2}P_{ab}^{2}\omega ^{2}}{\hbar ^{2}}\biggr)c_{b,n+1}\,.
557: \end{eqnarray}%
558: We consider as the initial state the atom in the excited state with the
559: cavity with a field characterized by coefficients $c_{n}(0)$, so that the
560: results are
561: \begin{eqnarray}
562: c_{a,n} &\simeq &c_{a,n}(0)\cos \biggl[\zeta _{n}\sqrt{n+1}gt\biggr]\,,
563: \label{CAn_sol} \\
564: c_{b,n+1} &\simeq &-ic_{a,n}(0)\sin \biggl[\zeta _{n}\sqrt{n+1}gt\biggr]\,.
565: \label{CBn_sol}
566: \end{eqnarray}%
567: with
568:
569: \begin{eqnarray}
570: \zeta_n\equiv\sqrt{1+\frac1{n+1}\biggl(\frac{ \text{v}_{x}P_{ab}\omega}{%
571: \hbar g}\biggr)^2}
572: \end{eqnarray}
573:
574: The average number of photons is
575: \begin{equation*}
576: \bar{n}(t)=\sum_{0}^{N}(nP_{n}(t)).
577: \end{equation*}%
578: where $P_{n}(t)=|c_{a,n}|^{2}+|c_{b,n}|^{2}$ is the probability for finding $%
579: n$ photons in the cavity. Now note from Eqs. (\ref{CAn_sol})-(\ref{CBn_sol})
580: that $P_{n}(t)=P_{n}(0)$. This shows that the photon statistics and
581: consequently the average number of photons is not altered by the background v%
582: $_{x}$.
583:
584: Considering an initial vacuum state in a cavity at resonance ($\Delta =0$),
585: we attain the following population inversion function:
586: \begin{equation}
587: W(t)=\cos (2g\zeta _{0}t). \label{W_vac}
588: \end{equation}%
589: For a sufficiently small background $\left( {\text{v}_{x}P_{ab}\nu /\hbar
590: g\ll 1}\right) $, it is allowed to write
591: \begin{equation}
592: \zeta _{0}\cong 1+\frac{1}{2}\frac{\text{v}_{x}^{2}P_{ab}^{2}\nu ^{2}}{\hbar
593: ^{2}g^{2}} \label{zeta_vac}
594: \end{equation}%
595: The former Eqs.(\ref{W_vac})-(\ref{zeta_vac}) mean that, at first
596: approximation, there appears an effective coupling (due to the background)
597: given by $g^{\prime }=g+{\text{v}_{x}^{2}P_{ab}^{2}\nu ^{2}}/({2\hbar ^{2}g}%
598: ) $. Consequently, this implies an increasing on the value of Rabi
599: frequency:
600: \begin{equation} \label{Omega_0}
601: \Omega _{0}=2g\biggl(1+\frac{1}{2}\frac{\text{v}_{x}^{2}P_{ab}^{2}\nu ^{2}}{%
602: \hbar ^{2}g^{2}}\biggr)
603: \end{equation}%
604: Regarding that discrepancies of 1 part in $10^{10}$ from the usual results
605: of quantum mechanics can be detected, we can limit the LV effects in
606: accordance with this sensitivity. We then impose that the correction term
607: should be smaller than $10^{-10}$, that is (v$_{x}P_{ab}\nu /\hbar
608: g)<10^{-10}.$ For the chosen parameters\textbf{\ }$P_{ab}\omega /\hbar
609: =3.2\times 10^{37}kg^{-1}m^{-1}$ and $g=66kHz$,\textbf{\ s}uch condition
610: leads to the following upper bound on the LV background: \textbf{$\text{v}$}$%
611: _{x}<2.06\times 10^{-38}kgm/s$, or in natural units $\text{v}$$%
612: _{x}<10^{-10}eV$\textbf{. }
613:
614: The same effect can be studied for an initial coherent state in a cavity
615: with medium number of photons $\bar{n}$ and at resonance ($\Delta =0$). In
616: this case we attain the following population inversion function:
617: \begin{equation}
618: W(t)=\sum_{n=0}^{\infty }\frac{{\bar{n}}^{n}e^{-{\bar{n}}}}{n!}\cos \biggl[2%
619: \sqrt{n+1}g\sqrt{1+\frac{\text{v}_{x}^{2}P_{ab}^{2}\omega ^{2}}{(n+1)\hbar
620: ^{2}g^{2}}}t\biggr]. \label{Wt_lv}
621: \end{equation}%
622: This expression allows to infer that the net effect of the background on the
623: probability amplitudes is the increasing of the frequency of the collapses
624: and revivals of the population inversion. It is instructive to note how the
625: high frequency terms of opposite phases in Eqs. (\ref{eom3})-(\ref{eom4})
626: conspired to provide such a correction. Note also that for v$_{x}=0$ the
627: usual result for the inversion $W(t)$ - a superposition of frequencies $2%
628: \sqrt{n+1}g$ - is recovered for an initial coherent state (see Eq. (\ref%
629: {W_coh})). For large $\bar{n},$ the sum in Eq. (\ref{Wt_lv}) can be
630: simplified to produce $W(t)\sim \cos (\Omega _{\bar{n}}t)$, where
631: \begin{equation}
632: \Omega _{\bar{n}}=2\sqrt{\bar{n}}g\sqrt{1+\frac{\text{v}_{x}^{2}P_{ab}^{2}%
633: \omega ^{2}}{\bar{n}\hbar ^{2}g^{2}}}
634: \end{equation}%
635: is the Rabi frequency corrected by the Lorentz-violating background. This
636: expression reveals that the Rabi frequency increases with the background
637: magnitude. For small background we can also write
638: \begin{equation}
639: \Omega _{\bar{n}}\cong 2\sqrt{\bar{n}}g\biggl({1+\frac{1}{2}\frac{\text{v}%
640: _{x}^{2}P_{ab}^{2}\omega ^{2}}{\bar{n}\hbar ^{2}g^{2}}}\biggr) \label{O_lv}
641: \end{equation}
642:
643: An important point here is the appearance of the mean photon number factor ($%
644: \bar{n})$ in the correction term in comparison with the previous case of an
645: initial vacuum state in the cavity (see Eq. (\ref{Omega_0})). Indeed, the
646: larger the average number of photons for an initial coherent state in the
647: cavity, the lower is the correction the Rabi frequency induced by the
648: fixed background. Considering the same experimental sensitivity of 1 part in
649: $10^{10}$ in a measurement of the Rabi frequency, a larger bound (less
650: stringent) on $\text{v}_{x}$ may now be achieved. In fact, for a state with
651: large number of photons $\left( \bar{n}>>1\right) $, the allowed background
652: upper value is multiplied by $\sqrt{\bar{n}}$ ($\text{v}_{x}^{COH}=\sqrt{%
653: \bar{n}}\text{v}_{x}^{VAC}$\textbf{$).$} This means that cavity experiments
654: with smaller number of photons imply better bounds (more stringent) on the
655: background magnitude. In this sense, the best probes for determining LV\
656: deviations from usual quantum mechanics are really the vacuum states.
657:
658: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
659: \begin{figure}[th]
660: \begin{center}
661: \includegraphics[{angle=0,width=8.5cm}]{fg_fig1.eps}
662: \end{center}
663: \caption{Population inversion as a function of time $C+W(t)$, for $\bar{n}%
664: =25 $ and (a) $C=1,\protect\nu _{x}=0$ (first, lower curve), (b) $C=3,%
665: \protect\nu _{x}=5\times10^{-33}kgm/s$ ( $2.5\times 10^{-5}eV$ in natural
666: units - second curve), (c) $C=5,\protect\nu _{x}=1\times 10^{-32}kgm/s$ ( $%
667: 5\times10^{-5}eV$ in natural units - third curve) and (d) $C=7,\protect\nu %
668: _{x}=2\times10^{-32}kgm/s$ (or $\protect\nu _{x}=1\times 10^{-4}eV$ -
669: fourth, upper curve). The $C$ constants are included to make easier the
670: comparison.}
671: \label{fig_Wt}
672: \end{figure}
673: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
674: To study the intermediate scale where effects of collapses and revivals
675: appear and to see the influence of the extra factor depending on the
676: background v$_{x}$ on the expression of Eq. (\ref{Wt_lv}), we proceed with a
677: graphical analysis. In Fig. 1, we present a plot of $W(t)\times gt$. Such a
678: figure shows a sequence of four curves for an initial coherent state with $%
679: \bar{n}=25$ and coupling $g=66kHz$. The first lower curve shows the usual
680: sequence of collapses and revivals in the absence of Lorentz violation $%
681: \left( \text{v}_{x}=0\right) $. As the background increases, the collapses
682: and revivals tend to occur later. Note also that the sequence of collapses
683: and revivals is always destroyed after some time, larger for higher values
684: of the background. These observations indicate that a stronger background
685: favours the maintenance of the revival/collapse sequence for a greater time.
686:
687: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
688: \begin{figure}[th]
689: \begin{center}
690: \includegraphics[{angle=0,width=8.5cm}]{fg_fig2a.eps} %
691: \includegraphics[{angle=0,width=8.5cm}]{fg_fig2b.eps} %
692: \includegraphics[{angle=0,width=8.5cm}]{fg_fig2c.eps}
693: \end{center}
694: \caption{Difference $W-W_0$ between population inversion functions $W$ for $%
695: \text{v}_x$ and $W_0$ for $\text{v}_x=0$, for $\bar{n}=25 $ where (a) $%
696: \protect\nu _{x}=5\times10^{-37}kgm/s$ ( $2.5\times 10^{-9}eV$ in natural
697: units - upper curve), (b) $\protect\nu _{x}=5\times10^{-38}kgm/s$ ( $%
698: 2.5\times 10^{-10}eV$ in natural units - middle curve), (c) $\protect\nu %
699: _{x}=5\times10^{-39}kgm/s$ ( $2.5\times 10^{-11}eV$ in natural units - lower
700: curve)}
701: \label{fig_Wt}
702: \end{figure}
703: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
704:
705: An interesting issue is to verify if this kind of analysis is able to impose
706: an upper bound on the background magnitude, in agreement to the analytical
707: result of Eq. (\ref{O_lv}). A reasonable criterion consists in taking the
708: maximum background value that does not yield significative discrepancy (of 1
709: part in $10^{10}$) on the population inversion pattern taking as reference
710: the usual case (v$_{x}=0)$. In this way, we have chosen to plot in Fig. 2
711: the difference $W-W_{0}$ at the time scale $0<gt<50$, sufficient to reveal
712: the beginning of the second collapse in the usual case (see lower picture in
713: Fig. 1). The comparative results from Fig. 2 shows that $\text{v}%
714: _{x}<10^{-10}eV$ is an efficient condition in keeping the difference $W-W_{0}
715: $ below $10^{-10}$ for the first observed sequence of collapses and revivals
716: of the population inversion. This is in agreement with our previous estimate
717: for the bound for an initial vacuum state in the cavity.
718:
719: Fig. 3 shows the same kind of graphical analysis of Fig. 1 for an initial
720: coherent state with lower mean number of photons $\left( \bar{n}=5\right) $.
721: Here, the background effect of keeping the sequence of collapses and
722: revivals for longer times is more clearly depicted. Further, it is seen that
723: the collapse time and separation time (between the revivals) increase with
724: the background magnitude. However, even lower background
725: values already reveal larger separations between collapses and revivals (in
726: comparison with the case of Fig. 1).
727:
728: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
729: \begin{figure}[th]
730: \begin{center}
731: \includegraphics[{angle=0,width=8.5cm}]{fg_fig3.eps}
732: \end{center}
733: \caption{Population inversion as a function of time $C+W(t)$, for $\bar{n}=5$%
734: . Values for $C$ and ${\text v}_x$ are the same from Fig. 1.}
735: \label{fig_Wt2}
736: \end{figure}
737: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
738:
739: A further step is to use Eq. (\ref{Wt_lv}) to study the implied changes on
740: the known expressions for the times $t_{R}$ (period of Rabi oscillations), $%
741: t_{c}$ (collapse time) and $t_{r}$ (revival time), defined in Ref. \cite%
742: {scully}. As before, we consider $\bar{n}\gg 1$. To measure the influence of
743: the background, it will be used the dimensionless parameter $\alpha =$v$%
744: _{x}P_{ab}\omega /(\hbar g)$. The time period $t_{R}$ is given by the
745: inverse of Rabi frequency%
746: \begin{equation}
747: t_{R}\sim \frac{1}{\Omega _{\bar{n}}}=\frac{1}{\Omega _{\bar{n}}^{0}}\bigg(1-%
748: \frac{\alpha ^{2}}{\bar{n}}\biggr), \label{eqtR}
749: \end{equation}%
750: outcome obtained using Eq. (\ref{O_lv}), with $\Omega _{\bar{n}}^{0}=2g\sqrt{%
751: \bar{n}}$. It shows that $t_{R}$ is reduced by the presence of the
752: background. This effect can be seen verifying that all the upper curves of
753: Fig. 3 exhibit higher oscillation frequency in comparison with the
754: lower ones (corresponding to lower v$_{x}$ values).
755:
756: Another parameter is the time of collapse of oscillations $\left(
757: t_{c}\right) $, that is, the time in which the oscillations associated with
758: different values of $n$ become uncorrelated. For an initial coherent state
759: in the cavity with sufficiently large $\bar{n}$ (for photon number standard
760: deviation $\Delta n=\sqrt{\bar{n}}\ll \bar{n}$), we can estimate $t_{c}$ as
761: (see Ref. \cite{scully}): $(\Omega _{\bar{n}+\sqrt{\bar{n}}}-\Omega _{\bar{n}%
762: -\sqrt{\bar{n}}})t_{c}\sim 1$. After using Eq. (\ref{O_lv}), we write
763: \begin{equation}
764: t_{c}\sim t_{c}^{0}\left( 1+\frac{\alpha ^{2}}{2\bar{n}}\right) ,
765: \label{eqtc}
766: \end{equation}%
767: where $t_{c}^{0}=1/(2g)$ is the collapse time in the usual two-level quantum
768: system (without LV). This expression shows an enlargement of the collapse
769: time with the background magnitude with fixed $\bar{n}$. Indeed, the upper
770: curves of Fig. 1 show longer collapse times when compared with the lower
771: ones. Also, when comparing Figs. 1 and 3 for the same background
772: values, we verify that the collapse time decreases with an increasing value
773: of $\ \bar{n}$. This is consistent with Eq. (\ref{eqtc}). This effect is
774: similar to the\ collapse time behavior observed in the usual \ two-level
775: quantum system (without LV) at a non-resonant regime $\left( \Delta \neq
776: 0\right) ,$ which also diminishes with $\bar{n}$ (see ref. \cite{scully}).
777:
778: Finally, we regard the time of revival of oscillations, $\left( t_{r}\right)
779: $, given by the condition $(\Omega _{\bar{n}}-\Omega _{\bar{n}-1})t_{r}=2\pi
780: m,\,\,m=1,2,...$, as the time in which two oscillators with
781: neighboring photon numbers $n=\bar{n}-1$ and $n=\bar{n}$ acquire a $2m\pi $
782: phase difference \cite{yoo}. This gives
783: \begin{equation}
784: t_{r}\sim {t_{r}^{0}}\biggl(1+\frac{\alpha ^{2}}{2\bar{n}}\biggr)
785: \label{eqtr}
786: \end{equation}%
787: with $t_{r}^{0}=2\pi m\sqrt{\bar{n}}/g,$ $m=1,2,3,...$. This result turns
788: clear that for a fixed background the revivals take place at regular
789: intervals as in the usual two-level quantum theory. Also, such intervals are
790: augmented for increasing values of the background, becoming the revival
791: packages more distant (in time) from each other. This is compatible with the
792: behavior exhibited by the upper curves in Figs. 1 and 3.
793:
794: \section{Conclusion}
795:
796: In this work, we have considered the main consequences of the LV vector
797: coupling term $\left( v_{\mu }\overline{\psi }\gamma ^{\mu }\psi \right) $
798: on a quantized two-level atom coupled with a quantized electromagnetic field
799: in a cavity. We have written the nonrelativistic LV corrections in the
800: interaction picture and considered such contributions into the Schr\"{o}%
801: dinger equation in order to obtain the modified system of coupled
802: differential equations for the probability amplitude coefficients that
803: describe the atom-field wave function.\ Experimental values of the relevant
804: parameters of the model revealed that the term $\mathbf{A}\cdot \mathbf{v}$
805: is of lower magnitude when compared with the term $\mathbf{p}\cdot \mathbf{v}
806: $, which corroborates recent results from the semiclassical theory. The
807: Lorentz-violating Hamiltonian stemming from the quantized vector potential
808: has a ``non-conserving" energy character. This kind of non-conserving terms
809: are usually discarded in the rotating wave approximation, but here are the
810: ones that implied physical effects on the system.
811:
812: After decoupling the system of differential equations, the probability
813: amplitude coefficients fulfill typical harmonic oscillator equations, whose
814: solutions lead to modified expressions for the population inversion function
815: and for the Rabi frequency. These results allow to note that the photon
816: statistics and the average number of photons in the cavity are not changed.
817: On the other hand, the Rabi frequency increases with the background value,
818: which is associated with a decreasing in the Rabi period $\left(
819: t_{R}\right) $. At the same time, the modified population inversion function
820: revealed that the revivals tend to occur later as the background magnitude
821: increases. As a consequence of the alteration implied on the Rabi frequency,
822: the collapse $\left( t_{c}\right) $ and revival $\left( t_{r}\right) $ times
823: increase with an increasing background, so that the revival packages become
824: larger and more distant from each other. In order to keep these
825: modifications in an undetectable scale for the parameter ranges considered,
826: an upper bound $\left( \text{v}_{x}\leq 10^{-10}eV\right) $ for the
827: background was set up. Such bound is in agreement with a recent result
828: obtained by us from a semiclassical approach \cite{Semi}.
829:
830: Note also that the deduced expressions for the modified times $t_{R},t_{c}$
831: and $t_{r}$ are better verified for $\bar{n}>>\sqrt{\bar{n}}$. In our
832: example, we have chosen $\bar{n}=25$ and $\bar{n}=5$, which does not fulfill
833: this condition. However, these not so large $\bar{n}$ values have already
834: obeyed qualitatively the tendency stated by Eqs. (\ref{eqtR}), (\ref{eqtc})
835: and (\ref{eqtr}). Despite a choice of \ larger values of $\bar{n}$
836: has implied a longer sequence of collapses and revivals, our
837: analysis showed that smaller values of $\bar{n}$ are more efficient in the
838: task of establishing a more stringent upper bound on the background
839: magnitude.
840: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
841:
842: Some interesting issues may be regarded in forthcoming investigations, as
843: the examination of the spontaneous emission in the presence of Lorentz
844: violation, with the evaluation of LV corrections on the decaying rate.
845:
846: \section*{Acknowledgments}
847:
848: The authors thank CNPq, FAPEMA (brazilian agencies), and CNPq-MCT-CT-Energ
849: for financial support. They are also grateful to K. Furuya for Ref. \cite%
850: {rwa} and B. Baseia for relevant discussions.
851:
852: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
853:
854: \begin{thebibliography}{99}
855: \bibitem{Samuel} V. A. Kostelecky and S. Samuel, \textit{Phys. Rev. Lett}.
856: \textbf{63}, 224 (1989); \textit{Phys. Rev. Lett}. \textbf{66}, 1811 (1991);
857: \textit{Phys. Rev. }D\textbf{\ 39}, 683 (1989); \textit{Phys. Rev. }D\textbf{%
858: \ 40}, 1886 (1989), V. A. Kostelecky and R. Potting, \textit{Nucl. Phys.} B%
859: \textbf{\ 359}, 545 (1991); Phys. Lett. B \textbf{381}, 89 (1996); V. A.
860: Kostelecky and R. Potting, \textit{Phys. Rev. }D\textbf{\ 51}, 3923 (1995).
861:
862: \bibitem{Colladay} D. Colladay and V. A. Kostelecky, \textit{Phys. Rev. }D
863: \textbf{55}, 6760 (1997); D. Colladay and V. A. Kostelecky, \textit{Phys.
864: Rev. }D \textbf{58}, 116002 (1998).
865:
866: \bibitem{Photon} S.M. Carroll, G.B. Field and R. Jackiw, \textit{Phys. Rev. }%
867: D \textbf{41}, 1231 (1990); C. Adam and F. R. Klinkhamer, \textit{Nucl.
868: Phys. }B \textbf{607}, 247 (2001); \textit{Nucl. Phys. }B \textbf{657}, 214
869: (2003); Phys. Lett. B\textbf{\ 513}, 245 (2001); C. Kaufhold and F. R.
870: Klinkhamer, \textit{Nucl. Phys. }B \textbf{734}, 1 (2006); R. Montemayor and
871: L.F. Urrutia, Phys. Rev. D \textbf{72}, 045018 (2005); X. Xue and J. Wu,
872: Eur. Phys. J. C \textbf{48}, 257 (2006); A.A. Andrianov and R. Soldati,
873: \textit{Phys. Rev.} D\textbf{\ 51}, 5961 (1995); \textit{Phys. Lett. }B%
874: \textbf{\ 435}, 449 (1998); A.A. Andrianov, R. Soldati and L. Sorbo, \textit{%
875: Phys. Rev.} D\textbf{\ 59}, 025002 (1998); Q. G. Bailey and V. A.
876: Kostelecky, Phys. Rev. D \textbf{70}, 076006 (2004); M. Frank and I. Turan,
877: Phys. Rev. D \textbf{74}, 033016 (2006); ~B. Altschul, \textit{Phys. Rev.} D
878: \textbf{72}, 085003 (2005); R. Lehnert and R. Potting, \textit{Phys. Rev.
879: Lett}. \textbf{93}, 110402 (2004); R. Lehnert and R. Potting, \textit{Phys.
880: Rev. D} \textbf{70}, 125010 (2004); ~B. Altschul, ~\textit{Phys. Rev. Lett.}
881: \textbf{98}, 041603 (2007); Phys.Rev. D \textbf{72, } 085003 (2005); Phys.
882: Rev. D \textbf{73}, 036005 (2006); Phys. Rev. D \textbf{73}, 045004 (2006);
883: H. Belich, M. M. Ferreira Jr, J.A. Helayel-Neto, M. T. D. Orlando, Phys.
884: Rev. D \textbf{67}, 125011 (2003); -ibid, Phys. Rev. D \textbf{69}, 109903
885: (E) (2004); H. Belich, M. M. Ferreira Jr, J.A. Helayel-Neto, M. T. D.
886: Orlando, Phys. Rev. D \textbf{68}, 025005 (2003); A. P. B. Scarpelli, H.
887: Belich, J. L. Boldo, J.A. Helayel-Neto, \textit{Phys. Rev. }D \textbf{67},
888: 085021 (2003).
889:
890: \bibitem{Radiative} R. Jackiw and V. A. Kosteleck\'{y}, \textit{Phys. Rev.
891: Lett.} \textbf{82}, 3572 (1999); J. M. Chung and B. K. Chung \textit{Phys.
892: Rev. }D \textbf{63}, 105015 (2001); J.M. Chung, \textit{Phys.Rev}. D\textbf{%
893: \ 60}, 127901 (1999); G. Bonneau, \textit{Nucl.Phys.} B\textbf{\ 593, }398
894: (2001); M. Perez-Victoria, \textit{Phys. Rev. Lett}. \textbf{83}, 2518
895: (1999); M. Perez-Victoria,\textit{\ JHEP} \textbf{0104}, 032 (2001); O.A.
896: Battistel and G. Dallabona, \textit{Nucl. Phys. B} \textbf{610}, 316 (2001);
897: O.A. Battistel and G. Dallabona, \textit{J. Phys. G} \textbf{28}, L23
898: (2002); \textit{J. Phys. G} \textbf{27}, L53 (2001); A. P. B. Scarpelli, M.
899: Sampaio, M. C. Nemes, and B. Hiller, \textit{Phys. Rev.} D\textbf{\ 64},
900: 046013 (2001); T. Mariz, J.R. Nascimento, E. Passos, R.F. Ribeiro and F.A.
901: Brito, JHEP 0510 (2005) 019.
902:
903: \bibitem{Topological} H. Belich, T. Costa-Soares, M.M. Ferreira Jr., J. A.
904: Helay\"{e}l-Neto, M.T. D. Orlando, \textit{Phys. Lett. B }\textbf{639, }675
905: (2006); H. Belich, T. Costa-Soares, M.M. Ferreira Jr., J. A. Helay\"{e}%
906: l-Neto, \textit{Eur. Phys. J. C }\textbf{41, }421 (2005), M. Lubo, \textit{%
907: Phys. Rev.} D \textbf{71}, 047701 (2005); M.N. Barreto, D. Bazeia, and R.
908: Menezes, \textit{Phys. Rev. D} \textbf{73}, 065015 (2006); L. R. Ribeiro, E.
909: Passos, C. Furtado, J. R. Nascimento, arXiv:0710.5858.
910:
911: \bibitem{CPT} R. Bluhm, V.A. Kostelecky, and N. Russell, \textit{Phys. Rev.
912: Lett}. \textbf{79}, 1432 (1997); R. Bluhm, V.A. Kostelecky, and N. Russell,
913: \textit{Phys. Rev. }D \textbf{57}, 3932 (1998); R. Bluhm, V.A. Kostelecky,
914: C. D. Lane, and N. Russell, \textit{Phys. Rev. Lett}. \textbf{88}, 090801
915: (2002); R. Bluhm and V.A. Kostelecky, \textit{Phys. Rev. Lett}. \textbf{84},
916: 1381 (2000); R. Bluhm, R. Bluhm, V.A. Kostelecky, and C. D. Lane, \textit{%
917: Phys. Rev. Lett}. \textbf{84}, 1098 (2000); R. Bluhm, V.A. Kostelecky,N.
918: Russell, \textit{Phys. Rev. Lett}. \textbf{82}, 2254 (1999); V.A. Kostelecky
919: and C. D. Lane, \textit{Phys. Rev. }D \textbf{60}, 116010 (1999); \ C. D.
920: Lane, \textit{Phys. Rev.} D \textbf{72}, 016005 (2005); N. Russell, \textit{%
921: J. Mod. Optics} \textbf{54}, 2481 (2007).
922:
923: \bibitem{Hydrogen} V.A. Kostelecky and C. D. Lane, J. Math. Phys. 40, 6245
924: (1999); R. Lehnert, \textit{J. Math. Phys.} \textbf{45}, 3399 (2004); M.M.\
925: Ferreira Jr and F. M. O. Moucherek, Int. J. Mod. Phys. \textbf{A} 21, 6211\
926: (2006); Nucl. Phys. A \textbf{790}, 635 (2007); H. Belich, T. Costa-Soares,
927: M.M. Ferreira Jr., J. A. Helay\"{e}l-Neto, and F. M. O. Moucherek, \textit{%
928: Phys. Rev. D} \textbf{74}, 065009 (2006); G. M. Shore, Nucl. Phys. \textbf{B}
929: 717, 86 (2005); O. G. Kharlanov and V. Ch. Zhukovsky, J. Math. Phys. \textbf{%
930: 48}, 092302 (2007).
931:
932: \bibitem{general} A. P. B. Scarpelli and J. A. Helayel-Neto, Phys. Rev. D
933: \textbf{73}, 105020 (2006); N.M. Barraz, Jr., J.M. Fonseca, W.A. Moura-Melo,
934: and J.A. Helayel-Neto, \textit{Phys. Rev. D} \textbf{76}, 027701 (2007);
935: J.W. Moffat, \textit{Int. J. Mod. Phys. D }\textbf{12} 1279 (2003); F. W.
936: Stecker and S.T. Scully, Astropart. Phys. \textbf{23}, 203 (2005); D. L.
937: Anderson, M. Sher and I. Turan, Phys. Rev. D \textbf{70}, 016001 (2004); I.
938: Turan, Int. J. Mod. Phys. A 19, 5395 (2004); H.\ Belich \textit{et al}.,
939: Phys. Rev. D\textbf{\ 68,} 065030 (2003); E. O. Iltan, Eur. Phys. J. C
940: \textbf{40}, 269 (2005); Mod. Phys. Lett. \textbf{A}19, 327 (2004); JHEP
941: 0306 (2003) 016; M. B. Cantcheff, \textit{Eur.Phys. J. C} \textbf{46}, 247
942: (2006); M. Botta Cantcheff, C.F.L. Godinho, A.P. Ba\^{e}ta Scarpelli, J.A.
943: Helay\"{e}l-Neto, \textit{Phys. Rev. D} \textbf{68}, 065025 (2003); S. Chen,
944: B. Wang, R. Su, \textit{Class. Quant. Grav. }\textbf{23, }7581 (2006); L.P.
945: Colatto, A.L.A. Penna, W.C. Santos, \textit{Eur. Phys. J. C} \textbf{36, }79
946: (2004); Q.G. Bailey, A. Kostelecky, \textit{Phys. Rev. D} \textbf{74},
947: 045001 (2006); M. Frank, I. Turan, and I.Yurdusen, arXiv:0709.4276; E.
948: Passos, L. R. Ribeiro, C Furtado, J. R. Nascimento, arXiv:0802.2817.
949:
950: \bibitem{Optical} H. Muller, C. Braxmaier, S. Herrmann, A. Peters, and C. L%
951: \"{a}mmerzahl, \textit{Phys. Rev.} D\textbf{\ 67}, 056006 (2003); H. M\"{u}%
952: ller, S. Herrmann,A. Saenz, A. Peters, and C. L\"{a}mmerzahl, P\textit{hys.
953: Rev.} D \textbf{68, }116006 (2003); H. M\"{u}ller, \textit{Phys. Rev.} D%
954: \textbf{\ 71}, 045004 (2005); N. Russell, \textit{Phys. Scripta }\textbf{72}%
955: , C38 (2005); D. F. Phillips, M. A. Humphrey, E. M. Mattison, R. E. Stoner,
956: R. F C. Vessot, R. L. Walsworth, \textit{Phys. Rev. D} \textbf{63}, 111101
957: (R) (2001); D. Bear, R.E. Stoner, R.L. Walsworth, V. A. Kostelecky, C. D.
958: Lane, \textit{Phys. Rev. Lett.} \textbf{85}, 5038 (2000); Erratum-ibid. 89
959: (2002) 209902; M.A. Humphrey, D.F. Phillips, E.M.Mattison, R.F.C.Vessot,
960: R.E.Stoner, R.L.Walsworth, \textit{Phys. Rev. A} \textbf{68}, 063807 (2003).
961: \bibitem{Semi} M. M. Ferreira Jr., A. R. Gomes, R. C. C. Lopes, \textit{%
962: Phys. Rev. D }\textbf{76}\textit{, }105031 (2007).
963:
964: \bibitem{scully} M. O. Scully and M. S. Zubairy, \textit{Quantum Optics},
965: Cambridge Univ. Press, Cambridge, 1997.
966:
967: \bibitem{foot1} The interaction picture may be defined by the following
968: equations: $|\Phi _{I}(t)\rangle =e^{iH_{0S}t/\hbar }|\Phi _{S}(t)\rangle ,$
969: $\widehat{A}_{I}(t)=e^{iH_{0S}t/\hbar }\widehat{A}_{S}e^{-iH_{0S}t/\hbar },$
970: where $H_{0S}$ is the Hamiltonian $H_{0}$ in the Schr\"{o}dinger
971: representation, $\widehat{A}_{I},\widehat{A}_{S}$ are the same operator $%
972: \widehat{A}$ written in the interactions and Schr\"{o}dinger pictures. These
973: two definitions yield the equations: $i\hbar \frac{d}{dt}|\Phi
974: _{I}(t)\rangle =H_{1}|\Phi _{I}(t)\rangle ,$ $\frac{d}{dt}\widehat{A}_{I}(t)=%
975: \frac{1}{i\hslash }[\widehat{A}_{I},H_{0I}].$
976:
977: \bibitem{loudon} R. Loudon, \textit{The quantum theory of light}, second
978: edition, Oxford Univ. Press, New York, 1983.
979:
980: \bibitem{foot2} It corresponds to a $\Delta m=0$ transition of a real atom.
981: See Chap. 15 from L. Mandel and E. Wolf, \textit{Optical Coherence and
982: quantum optics}, Cambridge Univ. Press, Cambridge, 1995.
983:
984: \bibitem{walls} D. F. Walls, G. J. Milburn, \textit{Quantum Optics},
985: Sringer-Verlag, Berlin, 1994.
986:
987: \bibitem{rwa} A. P. S. de Moura, \textquotedblleft Effect of
988: counter-rotating terms in Jaynes-Cummings model", dissertation, Unicamp 1997
989: (in portuguese). For the values $g=66kHz$ and $\omega =51GHz$ considered
990: later in this work, this means $\bar{n}\ll 10^{12}$ for RWA to be valid.
991:
992: \bibitem{prl96} M. Brune, F. Schmidt-Kaler, A. Maali, J. Dreyer, E. Hagley,
993: J. M. Ramond, S. Haroche, Phys. Rev. Lett. 76, 1800 (1996).
994:
995: \bibitem{prl94} M. Brune, P. Nussenzveig, F. Schmidt-Kaler, F. Bernardot, A.
996: Maali, J. M. Raimond, S. Haroche, Phys. Rev. Lett. 72, 3339 (1994).
997:
998: \bibitem{mil94} P. W. Milonni, \textit{The quantum vacuum}, Academic Press,
999: London, 1994.
1000:
1001: \bibitem{cook} R. J. Cook, Phys. Rev. A\textbf{29}, 1583 (1984).
1002:
1003: \bibitem{yoo} Y.I. Yoo and J.H. Eberly, Phys. Rep. \textbf{118}, 239 (1985).
1004: \end{thebibliography}
1005:
1006: \widetext
1007:
1008: \end{document}
1009: