0810.4208/cs.tex
1: %\documentclass[twocolumn,pra,showpacs,nofootinbib]{revtex4}
2: 
3: %\documentclass[twocolumn,floatfix,pra,showpacs,nofootinbib,groupedaddress]{revtex4}
4: \documentclass[aps,pra,showpacs,twocolumn,floatfix,groupedaddress]{revtex4}
5: 
6: \usepackage{amssymb}
7: \usepackage{amsmath}
8: \usepackage{amsfonts}
9: \usepackage{dcolumn,graphicx}
10: \usepackage{epsfig}
11: 
12: \begin{document}
13: 
14: \title{ AC Stark shift of the Cs microwave atomic clock transitions }
15: 
16: \author{P. Rosenbusch}
17: \affiliation {LNE-SYRTE, Observatoire de Paris, 75014 Paris, France}
18: 
19: \author{S. Ghezali}
20: \affiliation {Department of Physics, University Saâd Dahlab, Blida,
21:  and Laboratory of Quantum Electronics, University of Sciences and Technology HOUARI
22:  BOUMEDIENE, Bab-Ezzouar, Alger, Algeria}
23: 
24: \author{V. A. Dzuba}
25: \affiliation {
26: School of Physics, University of New South Wales, Sydney,
27: 2052, Australia}
28: 
29: \author{V. V. Flambaum}
30: \affiliation{
31: School of Physics, University of New South Wales, Sydney,
32: 2052, Australia}
33: 
34: \author{K. Beloy}
35: \affiliation {
36: Department of Physics, University of Nevada, Reno, Nevada 89557}
37: 
38: \author{A. Derevianko}
39: \affiliation {
40: Department of Physics, University of Nevada, Reno, Nevada 89557}
41: 
42: \date{\today}
43: 
44: \begin{abstract}
45: 
46: % We carry out a search for the magic wavelength on the clock transition of the Cs
47: % atom.
48: 
49: We analyze the AC Stark shift of the Cs microwave atomic clock
50: transition theoretically and experimentally. Theoretical and
51: experimental data are in a good agreement with each other.  Results
52: indicate the absence of a magic wavelength at which there would be
53: no differential shift of the clock states having zero projections of
54: the total angular momentum.% on the quantizing magnetic field.
55: \end{abstract}
56: 
57: \pacs{06.30.Ft, 37.10.Jk, 31.15.A-}
58: 
59: % 06.30.Ft Time and frequency
60: % 37.10.Jk Atoms in optical lattices
61: % 31.15.A- <i>Ab initio</i> calculations
62: 
63: \maketitle
64: 
65: \section{Introduction}
66: The present definition of the unit of time, the second, is based on
67: the frequency of the microwave transition between two hyperfine
68: levels of the Cs atom. Recently,  it has been realized that the
69: accuracy and stability of atomic clocks can be substantially
70: improved by trapping atoms in optical lattices operated at a certain
71: ``magic'' wavelength~\cite{KatTakPal03,YeKimKat08}. At this magic
72: wavelength, both clock levels experience the same AC Stark shift;
73: the clock frequency becomes essentially independent on trapping
74: laser intensity.
75: 
76: This effect was
77: demonstrated~\cite{TakHonHig05,LeTBaiFou06,LudZelCam08etal} for
78: optical clocks using strontium atoms. An extension of this idea to
79: microwave clocks with alkali-metal atoms Rb and Cs was considered in
80: Ref.~\cite{ZhoCheChe05}. A multitude of magic wavelengths for the
81: hyperfine transition was identified. Unfortunately, detailed
82: analysis presented below shows the conclusions of that paper to be
83: erroneous: there is no magic wavelength for Cs, at least for clock
84: levels with zero projections of the total angular momentum $M_F$ on
85: the quantizing magnetic field. In a separate
86: paper~\cite{FlaDzuDer08} we analyze the case of circular light
87: polarization and $M_F \neq 0$ levels and demonstrate that the AC
88: shift can be eliminated by an appropriate ``magic angle'' choice of
89: the direction of the magnetic field with respect to the light
90: propagation.
91: 
92: This paper presents a detailed theoretical analysis of the frequency
93: shift of a microwave clock involving a hyperfine transition. The
94: analysis requires calculation of the differential polarizability
95: involving third-order expressions, quadratic in the field strength
96: and linear in the hyperfine interaction. Evaluation of the resulting
97: expressions is carried out using relativistic many-body theory. The
98: second part of the paper reports measurements of the clock shift at
99: two laser wavelengths. The results of the calculations are in a good
100: agreement with the experimental measurements.
101: 
102: \section{Theory}
103: 
104: Here we follow the formalism of the quasi-energy states reviewed in
105: the context of laser-atom interaction in Ref.~\cite{ManOvsRap86}. We
106: start with considering the AC Stark shift $\delta E^{\left[ 2\right]
107: }$ in the second order of perturbation theory (quadratic in the
108: electric field) and then extend the formalism to the higher-order AC
109: Stark shift $\delta E^{\left[ 2+1\right]  }$ which takes into
110: account the hyperfine interaction (HFI). The latter shift appears in
111: the third order of perturbation theory and is quadratic in the field
112: amplitude and linear in the HFI. An important part of the analysis
113: involves the tensorial expansion of the shifts in the scalar,
114: vector, and tensor parts.
115: 
116: We are interested in transitions between two hyperfine components of
117: the same electronic states. Below we employ the conventional
118: labeling scheme for the atomic eigenstates, $\left\vert n\left(
119: IJ\right)  FM_{F}\right\rangle $, where $I$ is the nuclear spin, $J$
120: is the electronic angular momentum, and $F$ is the total angular
121: momentum, $\mathbf{F}=\mathbf{J}+\mathbf{I}$. $M_{F}$ is the
122: projection of $F$ on the quantization axis and $n$ encompasses the
123: remaining quantum numbers. Since the clock transitions involve the
124: same electronic state, we will also use a shorthand notation
125: $|F,M_F\rangle$. For example, the Cs fountain clock involves
126: transitions between hyperfine levels   $|F=4,M_F=0\rangle$ and
127: $|F'=3,M_F'=0\rangle$ of the $6s_{1/2}$ ground electronic state.
128: 
129: Under the influence of the laser each clock level is perturbed.
130: The clock frequency is modified by the difference in the perturbed energies
131: \begin{eqnarray}
132: \delta\nu^{\mathrm{Stark}}\left(  \omega_{L}\right)
133: &=&\frac{1}{h}\left[
134: \delta E_{n\left(  IJ\right)  F^{\prime\prime}M_{F}^{\prime\prime}%
135: }^{\mathrm{Stark}}\left(  \omega_{L}\right)
136: \right.\nonumber\\&&\left.-\delta E_{n\left( IJ\right)
137: F^{\prime}M_{F}^{\prime}}^{\mathrm{Stark}}\left( \omega_{L}\right)
138: \right] \label{Eq:ClockShift}
139: \end{eqnarray}
140: At the ``magic frequency'', this AC Stark clock shift would vanish.
141: 
142: \subsection{Second-order dynamic response}
143: 
144: This section introduces notation and reviews derivation and
145: tensorial analysis of the conventional second-order dynamic Stark
146: shift. We demonstrate that for states of total electronic angular
147: momentum $J=1/2$ and for $M_{F}=0$ levels (or for linear
148: polarization) the second-order AC Stark shift of the clock
149: transition vanishes.
150: 
151: Consider a traveling electromagnetic wave of an arbitrary polarization,
152: \[
153: \mathcal{\vec{E}}=\frac{1}{2}\mathcal{E}_{L}\hat{\varepsilon}~e^{-i\left(
154: \omega t-kz\right)  }+c.c.~,
155: \]
156: with the complex polarization vector parameterized by an angle $\theta$%
157: 
158: \begin{equation}
159: \hat{\varepsilon}=\mathbf{\hat{e}}_{x}\cos\theta+i\mathbf{\hat{e}}_{y}%
160: \sin\theta. \label{Eq:polVec}%
161: \end{equation}
162: The parametric angle $\theta$ may be related to the degrees of
163: linear, $l=\cos2\theta$, and circular, $A=\sin2\theta$,
164: polarization. Notice that the quantizing axis \thinspace$z$ is
165: chosen along the propagation vector $\mathbf{\hat{k}}$\textbf{ }of
166: the laser. The field amplitude $\mathcal{E}_{L}$ is related to the
167: intensity of the laser as $I_{L}=\frac
168: {c}{8\pi}\mathcal{E}_{L}^{2}$, or in practical units $I_{L}\left[
169: \frac{\mathrm{mW}}{\mathrm{cm}^{2}}\right]  \approx1.33\times\left(
170: \mathcal{E}_{L}\left[  \frac{\mathrm{V}}{\mathrm{cm}}\right] \right)
171: ^{2}$.
172: 
173: In the dipole approximation, the coupling can be represented as ($h.c.$ is a
174: hermitian conjugate)%
175: \[
176: V_{E1}\left(  t\right)  \equiv-\mathcal{\vec{E}}\cdot\mathbf{D}=-\frac{1}%
177: {2}\mathcal{E}_{L}~\hat{\varepsilon}\cdot De^{- i\omega t}+h.c..
178: \]
179: Application of the Floquet formalism (dressed states) yields the second-order
180: AC shift of the atomic energy level $a$%
181: \begin{eqnarray*}
182: \delta E_{a}^{\left[  2\right]  }&=&\sum_{b}\frac{\left\vert
183: \langle\psi _{b}|v|\psi_{a}\rangle\right\vert ^{2}}{E_{a}-\left(
184: E_{b}-\omega\right)
185: }
186: %\\&&
187: +\sum_{b}\frac{\left\vert \langle\psi_{a}|v|\psi_{b}\rangle\right\vert ^{2}%
188: }{E_{a}-\left(  E_{b}+\omega\right)  },
189: \end{eqnarray*}
190: where
191: $v=-\frac{1}{2}\mathcal{E}_{L}~\hat{\varepsilon}\cdot\mathbf{D}\,$,
192: $\psi_{a}$ and $\psi_{b}$ being the stationary atomic states with
193: unperturbed energies $E_{a}$ and $E_{b}$, respectively.
194: 
195: Now we proceed to the conventional reduction of the polarizability
196: into a sum over irreducible tensor operators. Introducing the
197: resolvent operator
198: ($H_{0}$ is the unperturbed atomic Hamiltonian)%
199: \[
200: R_{E_{a}}\left(  \omega\right)  =\left(  E_{a}-\hat{H}_{0}+\omega\right)
201: ^{-1},
202: \]
203: we may recast the shift as an expectation value $\delta E_{a}^{\left(
204: 2\right)  }=\left(  \frac{1}{2}\mathcal{E}\right)  ^{2}\langle\psi_{a}|\hat
205: {O}_{E1}\left(  \omega\right)  |\psi_{a}\rangle$, with%
206: \begin{eqnarray*}
207: \hat{O}_{E1}\left(  \omega\right)&=&\left( ~\hat{\varepsilon}\cdot
208: \mathbf{D}\right)  ^{\dagger}R_{E_{a}}\left( \omega\right) \left(
209: ~\hat {\varepsilon}\cdot\mathbf{D}\right)\\
210: &&+\left( ~\hat{\varepsilon}\cdot\mathbf{D}\right) R_{E_{a}}\left(
211: -\omega\right)  ~\left(  \hat{\varepsilon}\cdot \mathbf{D}\right)
212: ^{\dagger}.
213: \end{eqnarray*}
214: The order of coupling of the operators may be changed%
215: \begin{eqnarray*}
216: \hat{O}_{E1}\left(  \omega\right) &=&\sum_{K=0,1,2}\left[\left(
217: -1\right) ^{K}\right.\\
218: &&\times\left\{
219: \hat{\varepsilon}^{\ast}\otimes~\hat{\varepsilon}\right\}
220: _{K}\cdot\left\{  \mathbf{D}\otimes~R_{E_{a}}\left(  \omega\right)
221: \mathbf{D}\right\}  _{K}  \\&&\left.+\left\{
222: \hat{\varepsilon}^{\ast}\otimes~\hat{\varepsilon}\right\}
223: _{K}\cdot\left\{  \mathbf{D}\otimes~R_{E_{a}}\left( -\omega\right)
224: \mathbf{D}\right\}  _{K}\right]~,
225: \end{eqnarray*}
226: leading to the conventional decomposition into the scalar ($K=0$),
227: vector ($K=1$), and tensor ($K=2$) terms. Here we employed $\left\{
228: \hat {\varepsilon}\otimes~\hat{\varepsilon}^{\ast}\right\}
229: _{KM}=\left( -1\right)  ^{K}\left\{
230: \hat{\varepsilon}^{\ast}\otimes~\hat{\varepsilon }\right\}_{KM}$ and
231: the fact that $\hat{\varepsilon}$ and $\mathbf{D}$ are rank 1
232: tensors. The $M_{K}$ component of the compound tensor of rank $K$
233: composed from components of the tensors $A_{K_1}$ and $B_{K_2}$ (of
234: rank ${K_1}$ and ${K_2}$) is
235: defined as $\left\{A_{K_1}\otimes B_{K_2}~\right\}  _{KM_{K}}%
236: =\sum_{M_{1}M_{2}}C_{K_{1}M_{1}K_{2}M_{2}}^{KM_{K}}A_{K_{1}M_{1}}B_{K_{2}%
237: M_{2}}$, where $C_{{K_1}M_{1}{K_2}M_{2}}^{KM_{K}}$ are the
238: Clebsch-Gordan coefficients. The generalized scalar product is
239: defined as $\left( A_{K}\cdot B_{K}\right)  =\sum_{M_{K}}\left(
240: -1\right) ^{M_{K}}A_{KM_{K}}B_{K,-M_{K}}$.
241: 
242: Using the Wigner-Eckart theorem, a matrix element between two atomic
243: states
244: may be expressed as%
245: \begin{eqnarray*}
246: \langle FM_{F}|\hat{O}_{E1}\left(  \omega\right)
247: |FM_{F}^{\prime}\rangle&=&\sum_{K=0,1,2}\left(  -1\right)
248: ^{K}\sum_{\mu}\left(  -1\right) ^{\mu }\\&&\times\left\{
249: \hat{\varepsilon}^{\ast}\otimes~\hat{\varepsilon}\right\}  _{K,-\mu
250: }   \left(  -1\right)  ^{F-M_F}\\
251: &&\times\left(
252: \begin{tabular}
253: [c]{lll}%
254: $F$ & $K$ & $F$\\
255: $-M_{F}$ & $\mu$ & $M_{F}^{\prime}$%
256: \end{tabular}
257: \right)  \alpha_{nF}^{\left(  K\right)  }\left(  \omega\right)  ,
258: \end{eqnarray*}
259: with the reduced polarizabilities
260: \begin{eqnarray}
261: \alpha_{nF}^{\left(  K\right)  }\left(  \omega\right)   &  =&\langle
262: nF||\left\{  D\otimes~R_{E_{nF}}\left(  \omega\right)  D\right\}
263: _{K}\nonumber\\&&+\left( -1\right)  ^{K}\left\{
264: D\otimes~R_{E_{nF}}\left( -\omega\right)  D\right\}
265: _{K}||nF\rangle\nonumber\\
266: &=& \sqrt{[K]}(-1)^{K+2F}\sum_{F^{\prime}}\left\{
267: \begin{array}
268: [c]{ccc}%
269: 1 & 1 & K\\
270: F & F & F^{\prime}%
271: \end{array}
272: \right\} \nonumber\\&&\times\sum_{n^{\prime}}    \langle
273: nF||D||n^{\prime}F^{\prime}\rangle\langle
274: n^{^{\prime}}F^{\prime}||D||nF\rangle\nonumber \\
275: &&
276: \times\left(  \frac{1}{E_{nF}-E_{n^{^{\prime}%
277: }F^{^{\prime}}}+\omega}\right.\nonumber\\&&\left.+\left(  -1\right)
278: ^{K}\frac{1}{E_{nF}-E_{n^{^{\prime
279: }}F^{^{\prime}}}-\omega}\right)  . \label{Eq:alphared}%
280: \end{eqnarray}
281: Here we have used the shorthand notation $[K]\equiv(2K+1)$. The
282: matrix element may be simplified further using specific
283: parametrization, Eq.~(\ref{Eq:polVec}), of the polarization vector.
284: Explicitly,$\left\{
285: \hat{\varepsilon}^{\ast}\otimes~\hat{\varepsilon}\right\}
286: _{00}=-\frac {1}{\sqrt{3}},$ $\left\{
287: \hat{\varepsilon}^{\ast}\otimes~\hat{\varepsilon }\right\}
288: _{1\mu}=-\frac{\sin2\theta}{\sqrt{2}}\delta_{\mu,0}~,\left\{
289: \hat{\varepsilon}^{\ast}\otimes~\hat{\varepsilon}\right\}
290: _{2\mu}=-\frac
291: {1}{\sqrt{6}}\delta_{\mu,0}+\frac{1}{2}\cos2\theta~\delta_{\mu,\pm2}.$
292: 
293: Finally, the  AC Stark energy shift reads%
294: \begin{eqnarray}
295: \delta E_{nFM_{F}}^{\left(  2\right)  }&=&-\left(  \frac{1}{2}\mathcal{E}%
296: _{L}\right)  ^{2}\left[  \alpha_{nF}^{s}\left(  \omega\right)
297: +A~\alpha _{nF}^{a}\left(  \omega\right)
298: \frac{M_{F}}{2F}\right.\nonumber\\&&\left.-\alpha_{nF}^{T}\left(
299: \omega\right) \frac{3M_{F}^{2}-F\left(  F+1\right)  }{2F\left(
300: 2F-1\right)
301: }\right]  ,\label{Eq:StarkEnergyShift}%
302: \end{eqnarray}
303: with the conventional scalar, vector, and tensor polarizabilities
304: \begin{align}
305: \alpha_{nF}^{s}\left(  \omega\right)   &  =\frac{1}{\sqrt{3}}\frac{1}%
306: {\sqrt{[F]}}\alpha_{nF}^{\left(  0\right)  }\left(  \omega\right)
307: ,\nonumber\\
308: \alpha_{nF}^{a}\left(  \omega\right)   &  =-\frac{1}{\sqrt{2}}\frac{1}%
309: {\sqrt{[F]}}\frac{2F}{\sqrt{F\left(  F+1\right)  }}\alpha_{nF}^{\left(
310: 1\right)  }\left(  \omega\right)  ,\label{Eq:TraditionalContribs}\\
311: \alpha_{nF}^{T}\left(  \omega\right)   &  =-\frac{2}{\sqrt{6}}2F\left(
312: 2F-1\right)  \left[  \frac{\left(  2F-2\right)  !}{\left(  2F+3\right)
313: !}\right]  ^{1/2}\alpha_{nF}^{\left(  2\right)  }\left(  \omega\right)
314: .\nonumber
315: \end{align}
316: In general, there is an off-diagonal $M_{F}=M_{F}^{\prime}\pm2$
317: optical coupling involving the tensor part of the polarizability. In
318: practice, a quantizing $\mathbf{B}$-field is applied along the
319: propagation of the laser wave, and as long as the off-diagonal
320: coupling is much smaller than the Zeeman intervals, it can be
321: disregarded.
322: 
323: Since the dipole matrix elements do not couple to the nuclear degrees of
324: freedom, the dependence of the reduced polarizabilities on $I$ and $F\,\ $may
325: be be factored out as
326: \[
327: \alpha_{nF}^{\left(  K\right)  }\left(  \omega\right)  =(-1)^{J+I+F+K}%
328: \;\left[  F\right]  \left\{
329: \begin{array}
330: [c]{ccc}%
331: J & F & I\\
332: F & J & K
333: \end{array}
334: \right\}  \bar{\alpha}_{nJ}^{\left(  K\right)  }\left(  \omega\right)  ,
335: \]
336: where the quantities$~\bar{\alpha}_{nJ}^{\left(  K\right)  }\left(
337: \omega\right)  $ are the reduced matrix elements in the $|nJM_{J}\rangle$
338: basis,
339: \begin{eqnarray*}
340: \bar{\alpha}_{nJ}^{\left(  K\right)  }\left(  \omega\right)
341: &=&\langle nJ||\left\{  D\otimes~R_{E_{nF}}\left(  \omega\right)
342: D\right\} _{K}\\&&+\left( -1\right)  ^{K}\left\{
343: D\otimes~R_{E_{nF}}\left( -\omega\right)  D\right\} _{K}||nJ\rangle.
344: \end{eqnarray*}
345: These quantities do not depend on either $I$ or $F$.
346: 
347: With this factorization, we can make important comments specific to
348: the case of $J=1/2$ (e.g., ground state of alkali-metal atoms such
349: as Rb and Cs). Due to the angular selection rules, the tensor
350: contribution (expectation value of the rank 2 tensor) vanishes and
351: the only contributions come from the scalar and vector parts. As the
352: vector part of the energy shift is proportional to $M_{F}$, for
353: $M_{F}=0$ clock levels only the scalar contribution remains. The
354: vector contribution also vanishes for the case of linear
355: polarization ($A=0$).
356: 
357: The next important step is to demonstrate that the scalar shift does not
358: depend on $F$. In other words, there is no clock shift at the second order.
359: Indeed, for the scalar term, $\alpha_{nF}^{s}\left(  \omega\right)  =\frac
360: {1}{\sqrt{3}}\frac{1}{\sqrt{[F]}}\alpha_{nF}^{\left(  0\right)  }\left(
361: \omega\right)  =\frac{1}{\sqrt{3}}\frac{1}{\sqrt{[J]}}\bar{\alpha}%
362: _{nJ}^{\left(  0\right)  }\left(  \omega\right)  $. This result holds for an
363: arbitrary $J$.
364: 
365: This result has a very simple explanation. As it was pointed out
366: above, for the case of linear polarization of laser light and
367: $J=1/2$ we have only a scalar contribution to the polarizability.
368: This contribution does not depend on the orientation of the
369: quantization axis (external magnetic field). Let us consider the
370: case when the quantization axis is directed along the laser electric
371: field (along the linear polarization vector ${\bf e}$). From the
372: symmetry of the problem it is obvious that the electron states
373: $|J_z=1/2>$ and $|J_z=-1/2>$ have exactly equal quadratic shifts.
374: The hyperfine states $|F,F_z>$ are linear combinations of these
375: electron states multiplied by nuclear states. Since both components,
376: $|J_z=1/2>$ and $|J_z=-1/2>$, of any hyperfine state have the same
377: shift, all hyperfine states have the same shifts, i.e. the
378: differential polarizability is equal to zero. To have a non-zero
379: differential polarizability one has to include the hyperfine
380: interaction. Note that the inclusion of the magnetic polarizability
381: does not change this conclusion. If the lattice laser frequency is
382: in optical range, the magnetic polarizability contribution is
383: suppressed by an additional factor $\mu_B^2/D^2 \sim
384: \alpha^2\approx(1/137)^2$ and may be neglected.
385: 
386: To summarize, we  arrive at the conclusion that for $J=1/2$,
387: $M_{F}=0$ clock levels (or for linear polarization) the second-order
388: AC Stark shift is zero. Since the calculations of
389: Ref.~\cite{ZhoCheChe05} were limited to this second-order, their
390: conclusions are erroneous. This is also shown in the Appendix
391: without using irreducible tensor algebra.
392: 
393: %A similar consideration is valid for the magnetic polarizability shift
394: %which appears due to interaction of atomic magnetic moment with
395: %laser light magnetic field. However, this contribution to the shift
396: %is very small. Indeed, in magnetic polarizability terms there is frequency
397: %of laser light in denominator. It produces suppression of the order
398: %$\delta E_{\rm hfs}/\omega$.
399: 
400: \subsection{Non-trivial effect of the hyperfine interaction}
401: 
402: In the previous section, the HFS interaction has served the role of
403: an \textquotedblleft observer\textquotedblright, as it only defined
404: the coupling scheme. In particular, we find that the $M_{F}=0$
405: levels of the  hyperfine manifold attached to the $J=1/2$ levels are
406: shifted identically - at that level of approximation any laser
407: wavelength is \textquotedblleft magic\textquotedblright, i.e., the
408: clock transition remains unperturbed by any laser field. The
409: non-trivial effect arises when we take into account the dynamic (as
410: opposed to the observer) role of the HFS interaction.
411: 
412: Formally, this effect appears in the third-order double perturbation
413: theory with two laser and one HFS interactions. We build the
414: perturbation theory in terms of combined interaction
415: \[
416: V=V_{\mathrm{hfs}}+V_{E1}\left(  t\right)  .
417: \]
418: The convenience of the Floquet formalism is that we may immediately
419: employ the conventional formula for the third-order energy
420: correction
421: \begin{eqnarray*}
422: \delta E_{a}^{\left[  2+1\right]  }&=&\sum_{b,c\neq a}\frac{V_{ab}V_{bc}V_{ca}%
423: }{\left(  E_{b}^{\left(  0\right)  }-E_{a}^{\left(  0\right)  }\right)
424: \left(  E_{c}^{\left(  0\right)  }-E_{a}^{\left(  0\right)  }\right)  }%
425: \\&&-V_{aa}\sum_{b\neq a}\frac{V_{ab}V_{ba}}{\left(  E_{b}^{\left(
426: 0\right)
427: }-E_{a}^{\left(  0\right)  }\right)  ^{2}},%
428: \end{eqnarray*}
429: Here  $a,b,c$ are dressed atomic states, i.e.
430: $|a\rangle=|\psi_{a}\rangle e^{in\omega t}$, with $n$ representing
431: the number of photons ($n$ could be both negative and positive). The
432: scalar product, in addition to the conventional Hilbert space
433: operational definition includes an averaging over the period of
434: oscillation of the laser field. Explicitly, after the time averaging
435: (now $a,b,$ and $c$ are atomic states)
436: \[
437: \delta E_{a}^{\left[  2+1\right]  }\left(  \omega\right)
438: =T_{a}\left( \omega\right)  +C_{a}\left(  \omega\right)
439: +B_{a}\left(  \omega\right) +O_{a}\left(  \omega\right),
440: \]%
441: \begin{eqnarray*}
442: T_{a}\left(  \omega\right)   &  =&\langle
443: a|V_{\mathrm{hfs}}R_{E_{a}}\left( 0\right)  vR_{E_{a}}\left(
444: \omega\right)  v^{\dagger}|a\rangle\\&&+\langle
445: a|V_{\mathrm{hfs}}R_{E_{a}}\left(  0\right)
446: v^{\dagger}R_{E_{a}}\left(
447: -\omega\right)  v|a\rangle,\\
448: C_{a}\left(  \omega\right)   &  =&\langle a|vR_{E_{a}}\left(
449: \omega\right) V_{\mathrm{hfs}}R_{E_{a}}\left(  \omega\right)
450: v^{\dagger}|a\rangle\\&&+\langle a|v^{\dagger}R_{E_{a}}\left(
451: -\omega\right)  V_{\mathrm{hfs}}R_{E_{a}}\left(
452: -\omega\right)  v|a\rangle,\\
453: B_{a}\left(  \omega\right)   &  =&\left[  T_{a}\left(  \omega\right)
454: \right]
455: ^{\ast},\\
456: O_{a}\left(  \omega\right)   &  =&-\left(  V_{\mathrm{hfs}}\right)
457: _{aa}\left(  \langle a|v^{\dagger}\left(  R_{E_{a}}\left(
458: \omega\right) \right)  ^{2}v|a\rangle\right.\\&&\left.+\langle
459: a|v\left( R_{E_{a}}\left(  -\omega\right) \right)
460: ^{2}v^{\dagger}|a\rangle\right)  .
461: \end{eqnarray*}
462: Here $T_{a}\left(  \omega\right)  $, $C_{a}\left(  \omega\right)  $,
463: and $B_{a}\left(  \omega\right)  $ stand for top, center, and bottom
464: position of the HFS interaction in the respective diagram. The term
465: $O_{a}\left( \omega\right)  $ describes the normalization term. The
466: relevant diagrams are shown in Fig.~\ref{diagrams}.
467: 
468: \begin{figure}[h]
469: \centering
470: %\epsfig{figure=diagrams.eps,scale=0.5}
471: %\includegraphics*[scale=0.5]{diagrams-3rdOrder.eps}
472: %\includegraphics*[scale=0.5]{diagrams.eps}
473: \includegraphics*[scale=0.35]{fig1.eps}
474: \caption{Contributions to the dynamic polarizability
475: $\alpha(\omega)$. Diagram (a) represents the second-order
476: contribution arising from two photon interactions (wavy lines).
477: Diagrams (b-e) represent the additional effect of the hyperfine
478: interaction (capped line) and correspond respectively to the
479: third-order top, center, bottom, and normalization terms as
480: described in the text.} \label{diagrams}
481: \end{figure}
482: 
483: The interaction of the electron with the nuclear magnetic moment $\mu$ reads%
484: \[
485: V_{\mathrm{hfs}}=\left(  \mu\cdot\mathcal{T}^{\left(  1\right)  }\right)  ,
486: \]
487: where $\mathcal{T}^{\left(  1\right)  }$ is the rank 1 irreducible tensor
488: operator acting in the electronic coordinates with components
489: \[
490: \mathcal{T}_{\lambda}^{\left(  1\right)  }=-\frac{\left\vert e\right\vert
491: }{4\pi\varepsilon_{0}}\frac{i\sqrt{2}\left(  \mathbf{\alpha}\cdot
492: \mathbf{C}_{1\lambda}^{\left(  0\right)  }\left(  \mathbf{\hat{r}}\right)
493: \right)  }{cr^{2}},
494: \]
495: where $\mathbf{\alpha}$ stands for the  Dirac matrices and
496: $\mathbf{C}_{1\lambda}^{\left(  0\right)  }\left(
497: \mathbf{\hat{r}}\right)  $ are the normalized vector spherical
498: harmonics. In the formulas below we require the reduced matrix
499: element of the nuclear moment operator in the nuclear basis,
500: $\langle I||\mu||I\rangle$. It is related to the nuclear magnetic
501: $g$-factor as
502: \[
503: \langle I||\mu||I\rangle=\frac{1}{2}\sqrt{\left(  2I\right)  \left(
504: 2I+1\right)  \left(  2I+2\right)  }~g\mu_{n},
505: \]
506:  $\mu_{n}$ being the nuclear magneton.
507: 
508: Carrying out the angular decomposition similar to the second-order
509: analysis of the preceding section we find that  the expressions for
510: the shift, Eq.~(\ref{Eq:StarkEnergyShift}), remain the same with the
511: reduced polarizabilities $\alpha_{nF}^{\left(  K\right)  }\left(
512: \omega\right)  $ redefined as
513: \[
514: \alpha_{nF}^{\left(  K\right)  }\left(  \omega\right)  \rightarrow\alpha
515: _{nF}^{\left(  K\right)  }\left(  \omega\right)  +\beta_{nF}^{\left(
516: K\right)  }\left(  \omega\right)  .
517: \]
518: The third-order rank $K=0,1,2$ corrections are given by the sum over
519: contributions from diagrams of Fig.~\ref{diagrams}.
520: \[
521: \beta_{nF}^{\left(  K\right)  }\left(  \omega\right)  =2\beta_{nF}^{\left(
522: K\right)  }\left(  \omega;\mathrm{T}\right)  +\beta_{nF}^{\left(  K\right)
523: }\left(  \omega;\mathrm{C}\right)  +\beta_{nF}^{\left(  K\right)  }\left(
524: \omega;\mathrm{O}\right)  .
525: \]
526: Explicitly,
527: \begin{widetext}
528: \begin{align*}
529: \beta_{nF}^{\left(  K\right)  }\left(  \omega;\mathrm{T}\right)
530: %&
531: =[F]\sqrt{[K]}
532: %\\&
533: \sum_{J_{a}J_{b}}\left(  -1\right)  ^{J+J_{a}}\left\{
534: \begin{tabular}
535: [c]{lll}%
536: $I$ & $I$ & $1$\\
537: $J_{a}$ & $J$ & $F$%
538: \end{tabular}
539: \ \right\}  \left\{
540: \begin{tabular}
541: [c]{lll}%
542: $J$ & $1$ & $J_{b}$\\
543: $1$ & $J_{a}$ & $K$%
544: \end{tabular}
545: \ \right\}  \left\{
546: \begin{tabular}
547: [c]{lll}%
548: $K$ & $J$ & $J_{a}$\\
549: $I$ & $F$ & $F$%
550: \end{tabular}
551: \ \ \right\}  T_{J_{a}J_{b}}^{\left(  K\right)  }\left(  nJ,\omega\right)  ,
552: \end{align*}%
553: \begin{align*}
554: \beta_{nF}^{\left(  K\right)  }\left(  \omega;\mathrm{C}\right)
555: =[F]\sqrt{[K]} &  \sum_{J_{a}J_{b}}\sum_{J_{i}}[J_{i}](-1)^{2J_{a}+J_{b}%
556: +J}
557: %\times\\&
558: \left\{
559: \begin{array}
560: [c]{ccc}%
561: J & J & J_{i}\\
562: I & I & 1\\
563: F & F & K
564: \end{array}
565: \right\}  \left\{
566: \begin{array}
567: [c]{ccc}%
568: J & J & J_{i}\\
569: J_{a} & J_{b} & 1\\
570: 1 & 1 & K
571: \end{array}
572: \right\}  C_{J_{a}J_{b}}^{\left(  K\right)  }\left(  nJ,\omega\right)  ,
573: \end{align*}%
574: \begin{align*}
575: \beta_{nF}^{\left(  K\right)  }\left(  \omega;\mathrm{O}\right)   &
576: =\left( -1\right)  ^{2J+1}\left[  F\right]  \sqrt{\left[  K\right]
577: }\left\{
578: \begin{array}
579: [c]{ccc}%
580: 1 & I & I\\
581: F & J & J
582: \end{array}
583: \right\}
584: %\times\\&
585: \left\{
586: \begin{array}
587: [c]{ccc}%
588: J & J & K\\
589: F & F & I
590: \end{array}
591: \right\}  \sum_{J_{a}}\left\{
592: \begin{array}
593: [c]{ccc}%
594: K & J & J\\
595: J_{a} & 1 & 1
596: \end{array}
597: \right\}
598: %\times
599: O_{J_{a}}^{\left(  K\right)  }\left(  nJ,\omega\right)  .
600: \end{align*}
601: Here we introduced the reduced sums
602: \begin{eqnarray}
603: T_{J_{a}J_{b}}^{\left(  K\right)  }\left(  nJ,\omega\right)
604: &=&\langle
605: I||\mu||I\rangle\sum_{n_{b}}\sum_{n_{a}\neq n}   \langle nJ||\mathcal{T}%
606: ^{\left(  1\right)  }||n_{a}J_{a}\rangle\langle n_{a}J_{a}||D||n_{b}%
607: J_{b}\rangle\langle n_{b}J_{b}||D||nJ\rangle \nonumber \\
608: && \times \left(  \frac{1}{E-E_{a}}\frac{1}{E-E_{b}+\omega}+\left(
609: -1\right) ^{K}\left(  \omega\rightarrow-\omega\right)  \right),
610: \label{Eq:TopSum}
611: \end{eqnarray}%
612: \begin{eqnarray}
613: C_{J_{a}J_{b}}^{\left(  K\right)  }\left(  nJ,\omega\right)
614: &=&\langle I||\mu||I\rangle\sum_{n_{a}n_{b}}\langle
615: nJ||D||n_{a}J_{a}\rangle   \langle n_{a}J_{a}||\mathcal{T}^{\left(
616: 1\right)  }||n_{b}J_{b}\rangle\langle n_{b}J_{b}||D||nJ\rangle
617: \nonumber \\&&\times \left(
618: \frac{1}{E-E_{a}+\omega}\frac{1}{E-E_{b}+\omega}+\left( -1\right)
619: ^{K}\left(  \omega\rightarrow-\omega\right)  \right),
620: \label{Eq:CenterSum}
621: \end{eqnarray}%
622: \begin{eqnarray}
623: O_{J_{a}}^{\left(  K\right)  }\left(  nJ,\omega\right) &=&\langle
624: I||\mu||I\rangle\langle nJ||\mathcal{T}^{\left(  1\right)
625: }||nJ\rangle\sum_{n_{a}}\langle nJ||D||n_{a}J_{a}\rangle\langle
626: n_{a}J_{a}||D||nJ\rangle
627: %\nonumber\\&&\times
628: \left( \frac {1}{\left(
629: E-E_{a}+\omega\right)  ^{2}}+\left( -1\right) ^{K}\left(
630: \omega\rightarrow-\omega\right)  \right). \label{Eq:NormSum}
631: \end{eqnarray}
632: \end{widetext}
633: Notice that the angular momenta of the intermediate states $J_{a}$
634: and $J_{b}$ are fixed.
635: 
636: \subsection{Numerical evaluation}
637: \label{calculations}
638: To perform the calculations we use an {\em ab initio} approach
639: which has been described in detail in Ref.~\cite{AngDzuFla06}.
640: In this approach high accuracy is attained by including
641: important many-body and relativistic effects.
642: 
643: Calculations start from the relativistic Hartree-Fock (RHF) method
644: in the $V^{N-1}$ approximation. This means that the initial RHF
645: procedure is done for a closed-shell atomic core with the valence
646: electron removed. After that, the states of the external electron
647: are calculated in the field of the frozen core. Correlations are
648: included by means of the correlation potential
649: method~\cite{DzuFlaSil87}. We use the all-order correlation
650: potential $\hat \Sigma$ which includes two classes of the
651: higher-order terms: screening of the Coulomb interaction and
652: hole-particle interaction (see, e.g.,~\cite{DzuFlaSus89} for
653: details).
654: 
655: 
656: To calculate $\hat \Sigma$ we need a complete set of single-electron
657: orbitals. We use the B-spline technique~\cite{JohBluSap88} to
658: construct the basis. The orbitals are built as linear combinations of
659: 40 B-splines of order 9 in a cavity of radius 40$a_B$.
660: The coefficients are chosen from the condition that the
661: orbitals are the eigenstates of the RHF Hamiltonian $\hat H_0$ of the
662: closed-shell core. The $\hat \Sigma$ operator is
663: calculated with the technique which combines solving equations for
664: the Green functions (for the direct diagram) with the summation over
665: complete set of states (exchange diagram)~\cite{DzuFlaSus89}.
666: 
667: 
668: The correlation potential $\hat \Sigma$ is then used to build a new
669: set of single-electron states, the so-called Brueckner orbitals.
670: This set is to be used in the summation in equations
671: (\ref{Eq:TopSum}), (\ref{Eq:CenterSum}), and (\ref{Eq:NormSum}).
672: Here again we use the B-spline technique to build the basis. The
673: procedure is very similar to the construction of the RHF B-spline
674: basis. The only difference is that the new orbitals are now the
675: eigenstates of the $\hat H_0 + \hat \Sigma$ Hamiltonian.
676: 
677: Brueckner orbitals which correspond to the lowest valence states are
678: good approximations to the real physical states. Their quality can
679: be tested by comparing experimental and theoretical energies.
680: Moreover, their quality can be further improved by rescaling the
681: correlation potential $\hat \Sigma$ to fit the experimental energies
682: exactly. We do this by replacing the $\hat H_0 + \hat \Sigma$
683: Hamiltonian with $\hat H_0 + \lambda \hat \Sigma$, in which the
684: rescaling parameter $\lambda$ is chosen for each partial wave to fit
685: the energy of the first valence state. The values of $\lambda$ are
686: $\lambda_s=1$, $\lambda_p=0.97$, and $\lambda_d=0.95$. Note that
687: these values are very close to unity. This means that even without
688: rescaling the accuracy is good and only a small adjustment of $\hat
689: \Sigma$ is needed. Note also that since the rescaling procedure
690: affects not only energies but also the wave functions, it usually
691: leads to improved values of the matrix elements of external fields.
692: In fact, this is a semi-empirical method to include omitted
693: higher-order correlation corrections.
694: 
695: Matrix elements of the HFS and electric dipole operators are found
696: by means of the time-dependent Hartree-Fock (TDHF)
697: method~\cite{DzuFlaSil87,DzuFlaSus84}. This method is equivalent to
698: the well-known random-phase approximation (RPA). In the TDHF method,
699: the single-electron wave functions are presented in the form $\psi =
700: \psi_0 + \delta \psi$, where $\psi_0$ is the unperturbed wave
701: function. It is an eigenstate of the RHF Hamiltonian $\hat H_0$:
702: $(\hat H_0 -\epsilon_0)\psi_0 = 0$.  $\delta \psi$ is the correction
703: due to an external field. It can be found by solving the TDHF
704: equation
705: \begin{equation}
706:     (\hat H_0 -\epsilon_0)\delta \psi = -\delta\epsilon \psi_0 - \hat F \psi_0 -
707:   \delta \hat V^{N-1} \psi_0,
708:   \label{TDHF}
709: \end{equation}
710: where $\delta\epsilon$ is the correction to the energy due to the
711: external field ($\delta\epsilon\equiv 0$ for the electric dipole
712: operator), $\hat F$ is the operator of the external field
713: ($V_\mathrm{hfs}$ or $-\mathbf{D}\cdot \mathcal{E}$), and $\delta
714: \hat V^{N-1}$ is the correction to the self-consistent potential of
715: the core due to the external field.
716: 
717: The TDHF equations are solved self-consistently for all states in the core. Then the
718: matrix elements between any (core or valence) states $n$ and $m$ are given by
719: \begin{equation}
720:     \langle \psi_n | \hat F + \delta \hat V^{N-1} | \psi_m \rangle.
721:     \label{mel}
722: \end{equation}
723: The best results are achieved when $\psi_n$ and $\psi_m$ are the
724: Brueckner orbitals computed with rescaled correlation potential
725: $\lambda\hat{\Sigma}$.
726: 
727: We use equation (\ref{mel}) for all HFS and electric dipole matrix
728: elements in evaluating the top, center, bottom, and normalization
729: diagrams
730: (Eqs.~(\ref{Eq:TopSum}),(\ref{Eq:CenterSum}),(\ref{Eq:NormSum})),
731: except for the ground state HFS matrix element in the normalization
732: diagram where we use experimental data. The results are presented in
733: section \ref{results}.
734: 
735: \section{Experiment}
736: We measure the frequency shift of the Cs clock transition ($| F=3,
737: M_F=0 \rangle $ to $|{ F}=4, { M_F}=0 \rangle$) induced by a far
738: detuned laser beam. A Cs fountain clock \cite{Bize} is used. At each
739: cycle $\sim10^6$ atoms are loaded in an optical molasses and cooled
740: to below $2\,\mu$K. Moving molasses launches the atoms upwards with
741: a speed of $4.1$~m/s where they pass twice through a microwave
742: cavity thereby realizing Ramsey spectroscopy. The Zeeman degeneracy
743: is lifted by a 1.6~mG magnetic field aligned along the fountains
744: axis. The detuned laser beam is a traveling wave beam also aligned
745: on the axis of the fountain. The light polarization is linear with
746: respect to the light propagation. The beam waist is larger than the
747: 11~mm diameter opening in the microwave cavity. This assures that
748: all atoms passing through the opening and being detected experience
749: the light. The light intensity averaged over 1 cm$^2$ is measured by
750: a commercial powermeter (Newport 840-C) before entering the
751: fountains vacuum chamber. One intensity measurement is taken before
752: and one after each one-day run. The intensity drift between the two
753: is of the order of 1\%, however the error of the light intensity
754: experienced by each atom is rather high. This is due to our
755: ignorance of the exact intensity distribution, the exact atom
756: distribution and intensity losses in the vacuum window as well as
757: parasite reflections inside the vacuum chamber. We estimate the
758: intensity error as 20\%.
759: 
760: The frequency shift is measured by alternating the fountain's
761: configuration every 50 cycles. The first configuration is the
762: standard clock operation. The second configuration is identical to
763: the first plus the laser beam opened during the Ramsey period. This
764: assures that the atom preparation and cooling is not disturbed by
765: the  laser and that the atomic cloud is identical in the two
766: configurations. Hence, we can assume that all other clock shifts, in
767: particular collisions, are identical for the two configurations. The
768: absolute frequency is measured for each configuration against a
769: highly stable local oscillator exhibiting no significant drift
770: during several hundred cycles. The frequency shift induced by the
771: light calculates as the simple difference
772: \[
773: \delta \nu = \nu_{\rm with~light} - \nu_{\rm without~light}.
774: \]
775: The frequency shift is measured for a number of laser intensities.
776: Two sets of measurements are taken for light at wavelengths of 532
777: nm and 780 nm. The averages of each set weighted by the statistical
778: frequency uncertainty give a light shift of $(-3.51 \pm 0.7)\times
779: 10^{-4}$ Hz(W/cm$^2$)$^{-1}$ for 532 nm and $(-2.27 \pm 0.4)\times
780: 10^{-2}$ Hz(W/cm$^2$)$^{-1}$ for 780 nm. The statistical uncertainty
781: is negligible before the uncertainty on the light intensity.
782: % and relative statistical errors of $5\times10^{-2}$ and $4\times10^{-4}$,
783: %respectively.
784: 
785: 
786: \section{Results and discussion}
787: 
788: \label{results}
789: 
790: The shift of the clock frequency is given by (cf. Eq.~(\ref{Eq:ClockShift}))
791: \[
792: \delta \nu_L^\mathrm{Stark}=-\left(  \frac{1}{2}\mathcal{E}_L\right)  ^{2}\ \delta\alpha\left(
793: \omega_L\right),
794: \]
795: where $\delta\alpha\left(\omega_L\right)$ is the differential polarizability.
796: The conversion factor between differential polarizability in atomic units and the ratio of the shift to
797: laser intensity in practical units is given by
798: \[
799: \frac{\delta \nu_L^\mathrm{Stark}\left[  \mathrm{Hz}\right]  }{I_L\left[  \frac{\mathrm{mW}%
800: }{\mathrm{cm}^{2}}\right]  }=-4.68\times10^{-5}~\times~\delta\alpha\left(
801: \omega\right)  \left[  \mathrm{a.u.}\right] \, .
802: \]
803: We start by considering DC polarizabilities. In the static regime
804: ($\omega_L = 0$), our calculations give $\delta \alpha = 1.82 \times
805: 10^{-2}\,\mathrm{a.u.}$, which translates into the commonly used DC
806: Stark coefficient $k_S =-2.26 \times 10^{-10} \,
807: \mathrm{Hz}/(\mathrm{V/m})^2$. Notice that this value includes only
808: the scalar part of the polarizability. This is in agreement with the
809: most accurate experimental result~\cite{SimLauCla98} of  $k_S
810: =-2.271(4) \times 10^{-10} \, \mathrm{Hz}/(\mathrm{V/m})^2$.
811: 
812: \begin{figure}[ht]
813: \begin{center}
814: %\includegraphics*[scale=0.4]{cs0.eps}
815: \includegraphics*[scale=0.4]{fig2.eps}
816: %\epsfig[scale=0.4]{p0.eps}
817: \end{center}
818: \caption{(Color online) Differential polarizability of the Cs clock
819: transition ($M_F=M_F'=0$) in the
820: $\mathbf{B}\parallel\mathbf{\hat{k}}$ configuration as a function of
821: the probe laser frequency. Two experimental measurements (at 780 nm
822: and 532 nm) are compared with theoretical predictions (solid curve).
823: Refer to Fig.~\ref{Fig:ClockShift2} for better graphical resolution
824: of the 532 nm experimental point.} \label{Fig:ClockShift}
825: \end{figure}
826: 
827: For the AC case, our calculated differential polarizability $\delta
828: \alpha(\omega_L)$ for the cesium clock transition is presented in
829: Fig.~\ref{Fig:ClockShift} as a function of laser frequency. Both
830: values are given in atomic units. The two peaks correspond to the
831: $6s - 6p$ and $6s - 7p$ resonances. The graph never crosses zero,
832: which implies no magic frequency. Experimental results for two laser
833: wavelengths are also shown (also see Fig.~\ref{Fig:ClockShift2}).
834: Calculated and experimental relative frequency shifts are compared
835: in Table~\ref{tab:ex} and found to be in agreement with each other.
836: 
837: %Then for $\lambda=780~{\rm nm}$ ($\omega = 0.0584~{\rm a.u.}$) we get
838: %$s=-1.95\times10^{-2}~\mathrm{Hz}%
839: %/(\mathrm{W}/\mathrm{cm}^{2}\not )$ and for $\lambda=532~{\rm nm}$
840: %($\omega = 0.0856~{\rm a.u.}$) we get
841: %$s=-3.73\times10^{-4}~\mathrm{Hz}/(\mathrm{W}/\mathrm{cm}^{2})$. This is in
842: %good agreement with the experimental numbers:
843: % $-2.3\left(  4\right)  \times10^{-2}~\mathrm{Hz}%
844: %/(\mathrm{W}/\mathrm{cm}^{2}\not )$ and $-3.3\left(  6\right)  \times
845: %10^{-4}~\mathrm{Hz}/(\mathrm{W}/\mathrm{cm}^{2})$
846: 
847: 
848: \begin{figure}[ht]
849: \begin{center}
850: %\includegraphics*[scale=0.4]{cs1.eps}
851: \includegraphics*[scale=0.4]{fig3.eps}
852: %\epsfig[scale=0.4]{p0.eps}
853: \end{center}
854: \caption{(Color online) Same as in Fig.~\ref{Fig:ClockShift}, in the
855: region of the experimental point at 532 nm. }
856: \label{Fig:ClockShift2}
857: \end{figure}
858: 
859: \begin{table}[ht]
860: \caption{Comparison of the theoretical and experimental AC frequency
861: shifts for the clock transition in Cs} \label{tab:ex}
862: \begin{ruledtabular}
863: \begin{tabular}{cccc}
864:  $\lambda$,[nm] & $\omega$,[a.u.] & \multicolumn{2}{c}{$\delta \nu_L/I_L$, [Hz/mW/cm$^2$]} \\
865:                 &                 & Theor.  & Expt. \\
866: \hline
867: 780 & 0.0584 & $-1.95 \times 10^{-2}$ & $-2.27(40) \times 10^{-2}$ \\
868: 532 & 0.0856 & $-3.73 \times 10^{-4}$ & $-3.51(70) \times 10^{-4}$ \\
869: \end{tabular}
870: \end{ruledtabular}
871: \end{table}
872: 
873: 
874: To summarize, we presented a comprehensive analysis of the AC Stark
875: shift of the Cs microwave atomic clock transition. Theoretical
876: analysis based on the second and third order perturbation theory is
877: accompanied by measurements. Calculations and measurements are in
878: good agreement with each other and indicate the absence of a magic
879: frequency at least for the $M_F=0$ clock levels with zero
880: projections of the total angular momentum on the quantizing magnetic
881: field.
882: 
883: 
884: \appendix*
885: \section{}
886: Considering the complexities of working with the angular algebra,
887: here we analyze Eq.~(2) of Ref.~\cite{ZhoCheChe05} for the 2nd-order
888: dynamic Stark shift. Starting from their equation we again show that
889: there is no AC Stark shift of the clock transition frequency.
890: Eq.~(2) of Ref.~\cite{ZhoCheChe05} for the polarizability contains a
891: summation over $M^\prime$, which we manipulate
892: \begin{widetext}
893: \begin{eqnarray} \sum_{M^\prime}\left(
894: \begin{array}{ccc}
895:  F^\prime & 1 &  F \\
896:  M^\prime & p & -M
897: \end{array}
898: \right)^2&=&\left(-1\right)^{F^\prime-M+p}\sum_{M^\prime}\left(
899: \begin{array}{ccc}
900:  F &  1 &  F^\prime \\
901:  M & -p & -M^\prime
902: \end{array}
903: \right)\left(
904: \begin{array}{ccc}
905:  F^\prime & 1 &  F \\
906:  M^\prime & p & -M
907: \end{array}
908: \right)\nonumber\\
909: &=&\left(-1\right)^{F^\prime-M+p}\left(-1\right)^{2F}\sum_{KQ}
910: \left(-1\right)^{K-Q}\left[K\right]\left(
911: \begin{array}{ccc}
912:  F &  F &  K \\
913:  M & -M & -Q
914: \end{array}
915: \right)\left(
916: \begin{array}{ccc}
917:  K & 1 &  1 \\
918:  Q & p & -p
919: \end{array}
920: \right)\left\{
921: \begin{array}{ccc}
922:  1 & 1 & K \\
923:  F & F & F^\prime
924: \end{array}
925: \right\}\label{Eq:sumrule}\qquad\\
926: &=&\left(-1\right)^{F-M+p}\sum_{K}
927: \left(-1\right)^{F+F^\prime+K}\left[K\right]\left(
928: \begin{array}{ccc}
929:   F & K & F \\
930:  -M & 0 & M
931: \end{array}
932: \right)\left(
933: \begin{array}{ccc}
934:   1 & K & 1 \\
935:  -p & 0 & p
936: \end{array}
937: \right)\left\{
938: \begin{array}{ccc}
939:  1 & 1 & K \\
940:  F & F & F^\prime
941: \end{array}
942: \right\}\label{Eq:Q0}.
943: \end{eqnarray}
944: The expression~(\ref{Eq:sumrule}) is obtained from using the
945: summation rule 12.1(5) of Ref.~\cite{VarMosKhe88}; we further obtain
946: expression~(\ref{Eq:Q0}) by noting that only $Q=0$ terms are
947: non-zero in the summation.
948: 
949: Now we take the $F^\prime$-dependent part of~(\ref{Eq:Q0}) with the
950: $F^\prime$-dependent part of Eq.~(2) of Ref.~\cite{ZhoCheChe05} and
951: take the summation over $F^\prime$
952: \begin{eqnarray}
953: \sum_{F^\prime}\left(-1\right)^{F+F^\prime+K}\left[F^\prime\right]\left\{
954: \begin{array}{ccc}
955:  J        & J^\prime & 1 \\
956:  F^\prime & F        & I
957: \end{array}
958: \right\}^2\left\{
959: \begin{array}{ccc}
960:  1 & 1 & K \\
961:  F & F & F^\prime
962: \end{array}
963: \right\}&=&\sum_{F^\prime}\left(-1\right)^{F+F^\prime+K}\left[F^\prime\right]\left\{
964: \begin{array}{ccc}
965:  1 & F        & F^\prime \\
966:  I & J^\prime & J
967: \end{array}
968: \right\}\left\{
969: \begin{array}{ccc}
970:  I & J^\prime & F^\prime \\
971:  1 & F        & J
972: \end{array}
973: \right\}\left\{
974: \begin{array}{ccc}
975:  1 & F & F^\prime \\
976:  F & 1 & K
977: \end{array}
978: \right\}\nonumber\\
979: &=&\left(-1\right)^{I-J^\prime+F}\left\{
980: \begin{array}{ccc}
981:  J & J & K \\
982:  1 & 1 & J^\prime
983: \end{array}
984: \right\}\left\{
985: \begin{array}{ccc}
986:  J & J & K \\
987:  F & F & I
988: \end{array}
989: \right\}\label{Eq:sumrule2}.
990: \end{eqnarray}
991: The expression~(\ref{Eq:sumrule2}) is obtained from using the
992: summation rule 9.8(6) of Ref.~\cite{VarMosKhe88}.
993: 
994: Combining the above results gives
995: \begin{eqnarray}
996: \sum_{F^\prime M^\prime} \left[F\right]\left[F^\prime\right] \left(
997: \begin{array}{ccc}
998:  F^\prime & 1 &  F \\
999:  M^\prime & p & -M
1000: \end{array}
1001: \right)^2\left\{
1002: \begin{array}{ccc}
1003:  J        & J^\prime & 1 \\
1004:  F^\prime & F        & I
1005: \end{array}
1006: \right\}^2&=&\left(-1\right)^{F-M+p}\left(-1\right)^{I-J^\prime+F}
1007: \left[F\right]\sum_{K}\left[K\right]\left(
1008: \begin{array}{ccc}
1009:   F & K & F \\
1010:  -M & 0 & M
1011: \end{array}
1012: \right)\left(
1013: \begin{array}{ccc}
1014:   1 & K & 1 \\
1015:  -p & 0 & p
1016: \end{array}
1017: \right)\nonumber\\
1018: &&\times\left\{
1019: \begin{array}{ccc}
1020:  J & J & K \\
1021:  1 & 1 & J^\prime
1022: \end{array}
1023: \right\}\left\{
1024: \begin{array}{ccc}
1025:  J & J & K \\
1026:  F & F & I
1027: \end{array}
1028: \right\}.\nonumber
1029: \end{eqnarray}
1030: \end{widetext}
1031: Not surprisingly, the six-$j$ symbols here are identical to the ones
1032: appearing in the previously derived polarizabilities
1033: $\alpha^{(K)}_{nF}(\omega)$. Hence, this makes the connection to
1034: scalar ($K=0$), vector ($K=1$), and tensor ($K=2$) parts.
1035: 
1036: First we focus on the case $p=0$; this corresponds to linear
1037: polarization in the $\mathbf B
1038: \parallel \hat \varepsilon$ geometry. For $J=1/2$ atomic states the
1039: tensor part ($K=2$) is necessarily zero due to selection rules in
1040: the six-$j$ symbols (this is the case regardless of polarization).
1041: Furthermore the vector part ($K=1$) is zero due to the fact that the
1042: top row of the second three-$j$ symbol sums to an odd number (or see
1043: (\ref{Eq:3j}) below, with $p=0$). This leaves us with only the
1044: scalar part ($K=0$) to analyze. In this case, the r.h.s.~simply
1045: reduces to $1/(3[J])=1/6$. Thus we can conclude that for linear
1046: polarization of this geometry, the 2nd-order dynamic Stark shift is
1047: $F$-independent for $J=1/2$ atomic states.
1048: 
1049: 
1050: For the $\mathbf B
1051: \parallel \mathbf{\hat{k}}$ geometry, the linear polarization is regarded as
1052: an equal mixture of $\sigma^+$ ($p=+1$) and $\sigma^-$ ($p=-1$)
1053: circularly polarized light. Again the tensor part is necessarily
1054: zero for $J=1/2$. For the vector part, we note the three-$j$ symbol
1055: relation
1056: \begin{equation}
1057: \left(
1058: \begin{array}{ccc}
1059:   1 & K & 1 \\
1060:  -p & 0 & p
1061: \end{array}
1062: \right)=\left(-1\right)^K\left(
1063: \begin{array}{ccc}
1064:   1 & K &  1 \\
1065:   p & 0 & -p
1066: \end{array}
1067: \right).\label{Eq:3j}
1068: \end{equation}
1069: Thus, when we take equal mixtures of $\sigma^+$ and $\sigma^-$
1070: light, the vector contribution drops out. Again, we are left with
1071: only the scalar part. Not surprisingly, we again obtain the result
1072: $1/(3[J])=1/6$ when taking equal mixtures of $\sigma^+$ and
1073: $\sigma^-$ light.
1074: 
1075: 
1076: 
1077: The above results then generalize to any geometry for linearly
1078: polarized light.
1079: 
1080: 
1081: 
1082: \begin{acknowledgments}
1083: 
1084: This work was supported in part by the US National Science
1085: Foundation, by the Australian Research Council and by the US
1086: National Aeronautics and Space Administration under
1087: Grant/Cooperative Agreement No. NNX07AT65A issued by the Nevada NASA
1088: EPSCoR program. The SYRTE is Unit\'e de Recherche of the
1089: Observatoire de Paris and the Universit\'e Pierre et Marie Curie
1090: associated to the CNRS (UMR8630). It is a national metrology
1091: laboratory of the Laboratoire National de M\'etrologie et d'Essais
1092: (LNE) and member of the Institut francilien de recherche sur les
1093: atomes froids (IFRAF). S.G. acknowledges travel support from the
1094: LNE.
1095: \end{acknowledgments}
1096: 
1097: 
1098: 
1099: 
1100: 
1101: 
1102: 
1103: 
1104: 
1105: 
1106: 
1107: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1108: %reference to be added: Cold atom clocks and applications S Bize, P
1109: %Laurent, M Abgrall, H Marion, I Maksimovic, L Cacciapuoti, J
1110: %Grünert, C Vian, F Pereira dos Santos, P Rosenbusch, P Lemonde, G
1111: %Santarelli, P Wolf, A Clairon, A Luiten, M Tobar and C Salomon J.
1112: %Phys. B: At. Mol. Opt. Phys. 38 No 9 (14 May 2005) S449-S468
1113: 
1114: 
1115: %\bibliography{all}
1116: \begin{thebibliography}{16}
1117: \expandafter\ifx\csname
1118: natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1119: \expandafter\ifx\csname bibnamefont\endcsname\relax
1120:   \def\bibnamefont#1{#1}\fi
1121: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1122:   \def\bibfnamefont#1{#1}\fi
1123: \expandafter\ifx\csname citenamefont\endcsname\relax
1124:   \def\citenamefont#1{#1}\fi
1125: \expandafter\ifx\csname url\endcsname\relax
1126:   \def\url#1{\texttt{#1}}\fi
1127: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL
1128: }\fi \providecommand{\bibinfo}[2]{#2}
1129: \providecommand{\eprint}[2][]{\url{#2}}
1130: 
1131: \bibitem[{\citenamefont{Katori et~al.}(2003)\citenamefont{Katori, Takamoto,
1132:   Pal'chikov, and Ovsiannikov}}]{KatTakPal03}
1133: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Katori}},
1134:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Takamoto}},
1135:   \bibinfo{author}{\bibfnamefont{V.~G.} \bibnamefont{Pal'chikov}},
1136:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{V.~D.}
1137:   \bibnamefont{Ovsiannikov}}, \bibinfo{journal}{Phys.\ Rev.\ Lett.}
1138:   \textbf{\bibinfo{volume}{91}}, \bibinfo{pages}{173005}
1139:   (\bibinfo{year}{2003}).
1140: 
1141: \bibitem[{\citenamefont{Ye et~al.}(2008)\citenamefont{Ye, Kimble, and
1142:   Katori}}]{YeKimKat08}
1143: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Ye}},
1144:   \bibinfo{author}{\bibfnamefont{H.~J.} \bibnamefont{Kimble}},
1145:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Katori}},
1146:   \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{320}},
1147:   \bibinfo{pages}{1734} (\bibinfo{year}{2008}).
1148: 
1149: \bibitem[{\citenamefont{Takamoto et~al.}(2005)\citenamefont{Takamoto, Hong,
1150:   Higashi, and Katori}}]{TakHonHig05}
1151: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Takamoto}},
1152:   \bibinfo{author}{\bibfnamefont{F.~L.} \bibnamefont{Hong}},
1153:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Higashi}}, \bibnamefont{and}
1154:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Katori}},
1155:   \bibinfo{journal}{Nature (London)} \textbf{\bibinfo{volume}{435}},
1156:   \bibinfo{pages}{321} (\bibinfo{year}{2005}).
1157: 
1158: \bibitem[{\citenamefont{{Le Targat} et~al.}(2006)\citenamefont{{Le Targat},
1159:   Baillard, Fouch\'{e}, Brusch, Tcherbakoff, Rovera, and
1160:   Lemonde}}]{LeTBaiFou06}
1161: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{{Le Targat}}},
1162:   \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Baillard}},
1163:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fouch\'{e}}},
1164:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Brusch}},
1165:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Tcherbakoff}},
1166:   \bibinfo{author}{\bibfnamefont{G.~D.} \bibnamefont{Rovera}},
1167:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Lemonde}},
1168:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{97}},
1169:   \bibinfo{eid}{130801} (\bibinfo{year}{2006}).
1170: 
1171: \bibitem[{\citenamefont{Ludlow et~al.}(2008)\citenamefont{Ludlow, Zelevinsky,
1172:   and {Campbell {\em et al.}}}}]{LudZelCam08etal}
1173: \bibinfo{author}{\bibfnamefont{A.~D.} \bibnamefont{Ludlow}},
1174:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Zelevinsky}},
1175:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.~K.}
1176:   \bibnamefont{{Campbell {\em et al.}}}}, \bibinfo{journal}{Science}
1177:   \textbf{\bibinfo{volume}{319}}, \bibinfo{pages}{1805} (\bibinfo{year}{2008}).
1178: 
1179: \bibitem[{\citenamefont{Zhou et~al.}(2005)\citenamefont{Zhou, Chen, and
1180:   Chen}}]{ZhoCheChe05}
1181: \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Zhou}},
1182:   \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Chen}}, \bibnamefont{and}
1183:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Chen}}
1184:   (\bibinfo{year}{2005}), \eprint{arXiv:0512244}.
1185: 
1186: \bibitem[{\citenamefont{Flambaum et~al.}(2008)\citenamefont{Flambaum, Dzuba,
1187:   and Derevianko}}]{FlaDzuDer08}
1188: \bibinfo{author}{\bibfnamefont{V.~V.} \bibnamefont{Flambaum}},
1189:   \bibinfo{author}{\bibfnamefont{V.~A.} \bibnamefont{Dzuba}}, \bibnamefont{and}
1190:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Derevianko}}
1191:   (\bibinfo{year}{2008}), \eprint{arXiv:0809.2825}.
1192: 
1193: \bibitem[{\citenamefont{Manakov et~al.}(1986)\citenamefont{Manakov,
1194:   Ovsiannikov, and Rapoport}}]{ManOvsRap86}
1195: \bibinfo{author}{\bibfnamefont{N.~L.} \bibnamefont{Manakov}},
1196:   \bibinfo{author}{\bibfnamefont{V.~D.} \bibnamefont{Ovsiannikov}},
1197:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{L.~P.}
1198:   \bibnamefont{Rapoport}}, \bibinfo{journal}{Phys. Rep.}
1199:   \textbf{\bibinfo{volume}{141}}, \bibinfo{pages}{319} (\bibinfo{year}{1986}).
1200: 
1201: \bibitem[{\citenamefont{Angstmann et~al.}(2006)\citenamefont{Angstmann, Dzuba,
1202:   and Flambaum}}]{AngDzuFla06}
1203: \bibinfo{author}{\bibfnamefont{E.~J.} \bibnamefont{Angstmann}},
1204:   \bibinfo{author}{\bibfnamefont{V.~A.} \bibnamefont{Dzuba}}, \bibnamefont{and}
1205:   \bibinfo{author}{\bibfnamefont{V.~V.} \bibnamefont{Flambaum}}
1206:   (\bibinfo{year}{2006}), \bibinfo{note}{arXiv.org:physics/0605163}.
1207: 
1208: \bibitem[{\citenamefont{Dzuba et~al.}(1987)\citenamefont{Dzuba, Flambaum,
1209:   Silvestrov, and Sushkov}}]{DzuFlaSil87}
1210: \bibinfo{author}{\bibfnamefont{V.~A.} \bibnamefont{Dzuba}},
1211:   \bibinfo{author}{\bibfnamefont{V.~V.} \bibnamefont{Flambaum}},
1212:   \bibinfo{author}{\bibfnamefont{P.~G.} \bibnamefont{Silvestrov}},
1213:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{O.~P.}
1214:   \bibnamefont{Sushkov}}, \bibinfo{journal}{J. Phys. B}
1215:   \textbf{\bibinfo{volume}{20}}, \bibinfo{pages}{1399} (\bibinfo{year}{1987}).
1216: 
1217: \bibitem[{\citenamefont{Dzuba et~al.}(1989)\citenamefont{Dzuba, Flambaum, and
1218:   Sushkov}}]{DzuFlaSus89}
1219: \bibinfo{author}{\bibfnamefont{V.~A.} \bibnamefont{Dzuba}},
1220:   \bibinfo{author}{\bibfnamefont{V.~V.} \bibnamefont{Flambaum}},
1221:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{O.~P.}
1222:   \bibnamefont{Sushkov}}, \bibinfo{journal}{Phys.\ Lett.\ A}
1223:   \textbf{\bibinfo{volume}{141}}, \bibinfo{pages}{147} (\bibinfo{year}{1989}).
1224: 
1225: \bibitem[{\citenamefont{Johnson et~al.}(1988)\citenamefont{Johnson, Blundell,
1226:   and Sapirstein}}]{JohBluSap88}
1227: \bibinfo{author}{\bibfnamefont{W.~R.} \bibnamefont{Johnson}},
1228:   \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Blundell}},
1229:   \bibnamefont{and}
1230:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Sapirstein}},
1231:   \bibinfo{journal}{Phys.\ Rev.\ A} \textbf{\bibinfo{volume}{37}},
1232:   \bibinfo{pages}{307} (\bibinfo{year}{1988}).
1233: 
1234: \bibitem[{\citenamefont{Dzuba et~al.}(1984)\citenamefont{Dzuba, Flambaum, and
1235:   Sushkov}}]{DzuFlaSus84}
1236: \bibinfo{author}{\bibfnamefont{V.~A.} \bibnamefont{Dzuba}},
1237:   \bibinfo{author}{\bibfnamefont{V.~V.} \bibnamefont{Flambaum}},
1238:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{O.~P.}
1239:   \bibnamefont{Sushkov}}, \bibinfo{journal}{J. Phys. B}
1240:   \textbf{\bibinfo{volume}{17}}, \bibinfo{pages}{1953} (\bibinfo{year}{1984}).
1241: 
1242: \bibitem[{\citenamefont{Bize et~al.}(2005)\citenamefont{Bize, Laurent, Abgrall,
1243:   Marion, Maksimovic, Cacciapuoti, Grünert, Vian, dos Santos, Rosenbusch
1244:   et~al.}}]{Bize}
1245: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bize}},
1246:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Laurent}},
1247:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Abgrall}},
1248:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Marion}},
1249:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Maksimovic}},
1250:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Cacciapuoti}},
1251:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Grünert}},
1252:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Vian}},
1253:   \bibinfo{author}{\bibfnamefont{F.~P.} \bibnamefont{dos Santos}},
1254:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Rosenbusch}},
1255:   \bibnamefont{et~al.}, \bibinfo{journal}{J. Phys. B}
1256:   \textbf{\bibinfo{volume}{38}}, \bibinfo{pages}{S449} (\bibinfo{year}{2005}).
1257: 
1258: \bibitem[{\citenamefont{Simon et~al.}(1998)\citenamefont{Simon, Laurent, and
1259:   Clairon}}]{SimLauCla98}
1260: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Simon}},
1261:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Laurent}}, \bibnamefont{and}
1262:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Clairon}},
1263:   \bibinfo{journal}{Phys.\ Rev.\ A} \textbf{\bibinfo{volume}{57}},
1264:   \bibinfo{eid}{436-39} (\bibinfo{year}{1998}).
1265: 
1266: \bibitem[{\citenamefont{Varshalovich et~al.}(1988)\citenamefont{Varshalovich,
1267:   Moskalev, and Khersonskii}}]{VarMosKhe88}
1268: \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Varshalovich}},
1269:   \bibinfo{author}{\bibfnamefont{A.~N.} \bibnamefont{Moskalev}},
1270:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{V.~K.}
1271:   \bibnamefont{Khersonskii}}, \emph{\bibinfo{title}{Quantum Theory of Angular
1272:   Momentum}} (\bibinfo{publisher}{World Scientific},
1273:   \bibinfo{address}{Singapore}, \bibinfo{year}{1988}).
1274: 
1275: \end{thebibliography}
1276: 
1277: 
1278: \end{document}
1279: