hep-ph0111019/part1
1: \documentstyle[preprint,pra,aps]{revtex}
2: \begin{document}
3: \draft
4: \title{\bf Calculation of parity nonconservation in cesium
5: and possible deviation from the Standard Model}
6: \author{V.A. Dzuba, V.V. Flambaum, and J.S.M. Ginges}
7: \address{School of Physics, University of New South Wales,
8: Sydney 2052, Australia}
9: \date{\today}
10: \maketitle
11: 
12: \tightenlines
13: 
14: %************************************************************
15: \begin{abstract}
16: We have calculated the $6s-7s$ parity nonconserving E1 transition
17: amplitude ($E_{PNC}$) in cesium.
18: This calculation has been performed with higher numerical accuracy than
19: our 1989 calculation [V.A. Dzuba, V.V. Flambaum, and O.P. Sushkov,
20: Phys. Lett. A {\bf 141}, 147].
21: Also the Breit interaction has been included and the radiative
22: corrections estimated.
23: Our final result is $E_{PNC}=0.902\Big( 1\pm (0.7\%)\Big) iea_{B}(Q_{W}/N)$.
24: This represents an improvement in the accuracy of the calculation
25: from the $1\%$ error claimed in 1989.
26: This result corresponds to a nuclear weak charge for Cs,
27: $Q_{W}=-72.39\Big( 1\pm 0.4\%({\rm exp}) \pm 0.7\%({\rm theory})\Big)$.
28: We conclude that there is no significant deviation from the
29: Standard Model value $-73.09(3)$.
30: 
31: \end{abstract}
32: \vspace{1cm}
33: 
34: \pacs{PACS: 32.80.Ys, 11.30.Er, 12.15.Ji, 31.30.Jv}
35: %************************************************************
36: 
37: \section{Introduction}
38: 
39: Measurements of parity nonconservation (PNC) in atoms provide
40: a very useful test of the Standard Model.
41: In order to extract the value of the nuclear weak charge $Q_{W}$
42: from experiment, atomic structure calculations must be performed.
43: The most precise measurements and calculations have been performed
44: for cesium.
45: In 1989, calculations for cesium reached an accuracy of 1\%
46: \cite{dzuba89,blundell90}.
47: At this time, the experimental accuracy was at a level of 2\%.
48: In 1997, Wood {\it et al} measured the PNC amplitude in cesium
49: to an accuracy of 0.35\% \cite{wood97}.
50: In light of new measurements of quantities relevant to the PNC amplitude
51: (which are now in better agreement with the calculated values),
52: Bennett and Wieman analyzed the accuracy of the calculations
53: \cite{dzuba89,blundell90} and concluded that it is 0.4\% \cite{wieman99}.
54: With this accuracy, the nuclear weak charge deviates from the
55: Standard Model value by $2.5~\sigma$.
56: 
57: The suggestion that the actual accuracy of the calculations is $0.4\%$
58: immediately raised questions about values of small corrections which
59: were neglected in \cite{dzuba89,blundell90}.
60: It was shown by Derevianko \cite{derevianko00} that the contribution
61: of the Breit interaction to the PNC amplitude, $E_{PNC}$, is substantially
62: larger than previous estimates and reduces the deviation from the
63: Standard Model.
64: Sushkov pointed out \cite{sushkov01} that neglected radiative corrections
65: to $E_{PNC}$ must also be considered.
66: 
67: In this work we have performed a complete calculation of the $6s-7s$ PNC
68: amplitude in cesium.
69: We use the same method of calculation as that used
70: in the 1989 work \cite{dzuba89}
71: (this method is described in Section \ref{sec:method})
72: but with much better numerical accuracy.
73: The method which we call ``many-body perturbation theory in the screened
74: Coulomb interaction'' was developed to treat the most important
75: sequences of higher-order correlation diagrams in all orders.
76: Note that this is an all-order technique which is not a version of
77: the popular coupled-cluster technique.
78: This method has proven to be very accurate for alkaline
79: atoms and has produced very good results in a number of calculations.
80: We believe that it can hardly be improved in terms
81: of incorporating more correlation diagrams.
82: However, since more computer power is available to us now compared to
83: what we had in 1989 it is important to check the stability of the
84: results with the improved numerical accuracy.
85: In this work we have also calculated the Breit and radiative corrections
86: to the PNC amplitude. We have also accounted for the small shift in
87: $E_{PNC}$ due to consideration of the neutron distribution.
88: 
89: We list the results of our work below.
90: 
91: Repeating the 1989 calculation with much higher numerical accuracy,
92: we have obtained the same result:
93: $E_{PNC}=0.908$ (in units $iea_{B}(-Q_{W}/N)\times 10^{-11}$).
94: From the analysis below, we have concluded that the accuracy
95: of this value is about $0.5\%$.
96: 
97: Our calculation of the Breit interaction corrects $E_{PNC}$ by $-0.0055$.
98: This is in agreement with \cite{derevianko00,dzuba01,kozlov01}.
99: 
100: Radiative corrections to the weak charge
101: of order $\alpha$ have been calculated
102: in \cite{marciano} (see also \cite{lynnbed}).
103: However, there are important radiative corrections
104: of order $Z\alpha ^{2}$ and $Z^2\alpha ^{3} \ln^2(\lambda/R_n)$
105: which have recently been calculated in
106: \cite{milstein} (note that the latter correction is larger).
107: Here $\lambda$ is the electron Compton wavelength, and
108: $R$ is the nuclear radius.
109: Their contribution to $E_{PNC}$ is $0.004$. This contribution
110: originates from the radiative corrections to the weak matrix element
111: due to the Uehling potential (recently this contribution was calculated
112: in \cite{johnson01}).
113: We should add that there are other important contributions which have
114: the same magnitude.
115: In the sum-over-states approach (see Section \ref{sec:pnc})
116: they correspond to radiative corrections to the
117: energy intervals (-0.003) and E1 amplitudes (0.003).
118: Parametrically, they are proportional to
119: $Z^2\alpha ^{3} \ln(1/Z^2\alpha ^2)$. There are also corrections
120: to the weak matrix elements of order $Z^2\alpha ^{3} \ln(\lambda/R)$.
121: Our  estimate of these corrections shows that they have opposite
122: sign to the Uehling potential contribution and the same magnitude. As
123: a final result we suggest the following estimate for the contribution
124: of the radiative corrections to  $E_{PNC}$: $0.000 \pm 0.004$.
125: 
126: In this work we use an improved charge distribution
127: compared to that used in 1989.
128: This shifts $E_{PNC}$ very slightly, by 0.0007.
129: 
130: The neutron distribution in nuclei is slightly different
131: from the well-studied charge distribution. This produces
132: a small correction to $E_{PNC}$.
133: In this work we have found this correction to be $-0.0018$.
134: This is in agreement with the result of Ref. \cite{derevianko}.
135: 
136: This leaves us with $E_{PNC}=0.902 \pm 0.7 \%$.
137: Combined with measurements of $E_{PNC}/\beta$ from \cite{wood97}
138: and $\beta =27.00(6)a_{B}^{3}$ (the average value of the most accurate
139: results from \cite{dzuba00,wieman99} and \cite{dzuba97,wood97}) this gives
140: $Q_{W}=-72.39\Big( 1 \pm 0.4\%({\rm exp})\pm 0.7\%({\rm theory})\Big)$
141: which deviates by
142: $\Big( 1.0 \pm 0.4({\rm exp})\pm 0.7({\rm theory})\Big) \%$ from
143: the Standard Model value $Q_{W}= -73.09(3)$ \cite{groom00}.
144: From our point of view, this does not look like a significant
145: deviation from the Standard Model.
146: 
147: %************************************************************
148: \section{Method of calculation}
149: \label{sec:method}
150: 
151: 
152: The method that we use here was developed in the works
153: \cite{dzuba89,dzuba89energy,dzuba89e1hfs}.
154: As was emphasized in the introduction, it is an all-order technique,
155: in terms of treating correlations, and is not a version of the popular
156: coupled-cluster method.
157: The dominating sequences of higher-order correlation diagrams
158: correspond to real physical phenomena like screening of the
159: Coulomb interaction and the hole-particle interaction.
160: They are included in all orders in our technique.
161: While many-body perturbation theory in the
162: residual Coulomb interaction does not converge, there is good convergence
163: of our method ``many-body perturbation theory in the
164: screened Coulomb interaction''.
165: Another strong point of the method is that its complexity does not go
166: beyond the calculation of energies.
167: The most complicated and time consuming part of the method is the
168: calculation of the correlation potential $\hat{\Sigma}$.
169: $\hat{\Sigma}$ is used to calculate single-electron Brueckner orbitals.
170: The calculation of matrix elements with Brueckner orbitals
171: is easy and they already include most of the correlations.
172: This makes the calculation of hyperfine structure, PNC, etc. as simple as
173: the calculation of energies.
174: 
175: The nuclear spin-independent weak interaction of an electron with
176: the nucleus is
177: \begin{equation}
178: \label{eq:h_w}
179: \hat{H}_{W}=\frac{G_{F}}{2\sqrt{2}}\rho (r)Q_{W}\gamma _{5}
180: \end{equation}
181: where $G_{F}$ is the Fermi constant, $Q_{W}$ is the weak charge of the
182: nucleus, $\gamma _{5}$ is a Dirac matrix, and $\rho (r)$ is the nuclear
183: density (note that because the nuclear weak charge in the Standard
184: Model is approximately equal to the number of neutrons, the
185: density $\rho (r)$ should be taken as the neutron density;
186: see Section \ref{sec:pnc}).
187: 
188: There are essentially two different methods by which the PNC E1 amplitude
189: can be calculated:
190: from a ``mixed-states'' approach
191: (in which the external fields are taken into account
192: in the electron orbitals) and from a ``sum-over-states'' approach
193: (perturbation theory sum over intermediate opposite parity states).
194: The most complete calculation using the mixed-states method
195: was performed in the work \cite{dzuba89}.
196: In this work all-orders summation of dominating diagrams in
197: the residual electron-electron interaction was included.
198: This is the method used in the current work.
199: (For the sum-over-states method, see the discussion in Section
200: \ref{sec:pnc}.)
201: 
202: The method of calculation we use is applicable for $N$-electron atoms with
203: one valence electron. This method is particularly effective for cesium,
204: compared to other heavy atoms, as in this case the external electron has
205: very little overlap with the tightly-bound core, enabling the use of
206: perturbation theory in the calculation of the residual interaction of the
207: external electron with the core.
208: 
209: The calculations start from the relativistic Hartree-Fock (RHF) method
210: in the $\hat{V}^{N-1}$ approximation.
211: The single-electron RHF Hamiltonian is
212: \begin{equation}
213: \label{eq:RHF}
214: \hat{H}_{0}=c\mbox{\boldmath$\alpha$}\cdot \hat{{\bf p}}+(\beta -1)c^{2}-
215: Z\alpha/r+\hat{V}^{N-1} \ ,
216: \end{equation}
217: $\mbox{\boldmath$\alpha$}$ and $\beta$ are Dirac matrices and
218: $\hat{{\bf p}}$ is the electron momentum.
219: For cesium, the accuracy of the RHF energies is of the order of $10\%$.
220: 
221: We use the time-dependent Hartree-Fock (TDHF) method
222: (which is equivalent to the random-phase approximation with exchange)
223: to calculate the interaction of external fields with atomic electrons.
224: In this paper we deal with two external fields:
225: the electric field of the photon (E1 transition amplitudes)
226: and the weak field of the nucleus.
227: In the RHF approximation the interaction between an external field
228: $\hat{H}_{\rm ext}$ and atomic electrons is defined by the matrix element
229: $\langle \psi _{2}|\hat{H}_{\rm ext}|\psi _{1}\rangle$,
230: where $\psi _{1}$ and $\psi _{2}$ are RHF orbitals.
231: Inclusion of the polarization of the atomic core by an external field is
232: reduced to the addition of a correction $\delta \hat{V}$
233: (which is the correction to the Hartree-Fock potential due to the interaction
234: between the core and the external field)
235: to the operator which describes the interaction,
236: $\langle \psi _{2}|\hat{H}_{\rm ext}+
237: \delta \hat{V}|\psi _{1}\rangle$.
238: 
239: The TDHF contribution to $E_{PNC}$ between states $6s$ and $7s$ in
240: the mixed-states approach is given by
241: \begin{equation}
242: \label{eq:TDHF}
243: E_{PNC}^{TDHF}=\langle \psi _{7s}|\hat{H}_{E1}+\delta \hat{V}_{E1}|
244: \delta \psi _{6s}\rangle +
245: \langle \psi _{7s}|\hat{H}_{W}+\delta \hat{V}_{W}|
246: X _{6s}\rangle +
247: \langle \psi _{7s}|\delta \hat{V}_{E1W}|\psi _{6s}\rangle  \
248: \end{equation}
249: (of course, this term can instead be expressed in terms of corrections
250: to $\psi _{7s}$).
251: Here $\delta \psi $ and $\delta \hat{V}_{W}$ denote corrections to
252: single-electron RHF wavefunctions and the Hartree-Fock potential caused
253: by the weak interaction.
254: These corrections are found by solving
255: \begin{equation}
256: (\hat{H}_{0}-\epsilon)\delta \psi =
257: -(\hat{H}_{W}+\delta \hat{V}_{W})\psi \ .
258: \end{equation}
259: The positive (negative) frequency corrections $X$ ($Y$)
260: due to the E1 field of the photon are found from the equations
261: \begin{eqnarray}
262: (\hat{H}_{0}-\epsilon -\omega)X &=&
263: -(\hat{H}_{E1}+\delta \hat{V}_{E1})\psi \\
264: (\hat{H}_{0}-\epsilon +\omega)Y&=&
265: -(\hat{H}_{E1}^{\dagger}+\delta \hat{V}_{E1}^{\dagger})\psi  \ .
266: \end{eqnarray}
267: The correction to the potential due to the E1 field is
268: $\delta \hat{V}_{E1}$ ($\delta \hat{V}_{E1}^{\dagger}$).
269: The correction $\delta \hat{V}_{E1W}$ to the core potential
270: is due to the simultaneous action of the
271: weak field and the electric field of the photon.
272: 
273: The TDHF contribution to $E_{PNC}$ (Eq. \ref{eq:TDHF})
274: corresponds to calculating the lowest-order PNC diagrams
275: presented in Fig. \ref{fig:pnc} with the core polarization diagrams
276: (see Fig. \ref{fig:corepol}) included in all orders in the Coulomb
277: interaction.
278: 
279: The wavefunctions are improved by taking into account
280: electron-electron correlations.
281: The correlations are calculated using the ``correlation potential'' method
282: \cite{dzuba87} which corresponds to adding a non-local correlation potential
283: $\hat{\Sigma}$ to the potential $\hat{V}^{N-1}$ in the RHF Hamiltonian
284: (\ref{eq:RHF}) and then solving for the states of the external electron.
285: The correlation potential is defined such that its
286: average value coincides with the correlation correction to energy,
287: $\delta E_{i}=\langle\psi _{i}|\hat{\Sigma}|\psi _{i}\rangle$.
288: The correlation potential is calculated by means of many-body
289: perturbation theory in the residual Coulomb interaction
290: \begin{equation}
291: \hat{U}=\hat{H}-\sum_{i=1}^{N}\hat{H}_{0}({\bf r}_{i})=
292: \sum _{i<j}^{N}\frac{1}{|{\bf r}_{i}-{\bf r}_{j}|}-
293: \sum _{i=1}^{N}\hat{V}^{N-1}({\bf r}_{i}),
294: \end{equation}
295: where $\hat{H}$ is the exact Hamiltonian of an atom.
296: The lowest-order correlation diagrams $\hat{\Sigma}^{(2)}$
297: (second-order in $\hat{U}$) are presented in Fig.~\ref{fig:2ndorder}.
298: 
299: Brueckner orbitals are obtained by adding $\hat{\Sigma}^{(2)}$ to the
300: Hartree-Fock potential $\hat{V}^{N-1}$ and solving the equation
301: \begin{equation}
302: (\hat{H}_{0}+\hat{\Sigma}^{(2)}-\epsilon)\psi =0
303: \end{equation}
304: iteratively for the states of the external electron.
305: This corresponds to chaining the self-energy operator to all orders
306: (Fig.~\ref{fig:sechain2}).
307: This chain of diagrams is enhanced by the small denominator,
308: corresponding to the excitation energy of an external electron
309: (in comparison to the excitation of a core electron).
310: At this level of calculation (with ``bare'' $\hat{\Sigma}$)
311: the accuracy for energy levels for Cs is about 1\%.
312: 
313: Using the correlation potential method and the Feynman diagram technique
314: we include into $\hat{\Sigma} ^{(2)}$ two series of dominating higher
315: order diagrams which are calculated in all orders of perturbation theory
316: \cite{dzuba89,dzuba89energy,dzuba89e1hfs}.
317: These are screening of the electron-electron interaction
318: (Fig.~\ref{fig:screening}) and the hole-particle interaction
319: (Fig.~\ref{fig:hpchain}).
320: The electron-electron screening is a collective phenomenon.
321: The corresponding chain of diagrams is enhanced by a factor
322: approximately equal to the number of electrons in the external
323: closed subshell (the $5p$ electrons in Cs).
324: The hole-particle interaction is enhanced by the large
325: zero multipolarity diagonal matrix elements of the
326: Coulomb interaction.
327: This interaction accounts for the discrete spectra in noble gases.
328: We will denote the dressed self-energy operator
329: by $\hat{\Sigma}$ (Fig.~\ref{fig:hpscreense}).
330: With this $\hat{\Sigma}$ the Brueckner energies for Cs have an accuracy
331: of the order of $0.1\%$.
332: 
333: The wavefunctions can be further modified by placing a coefficient before
334: $\hat{\Sigma}$ such that the corresponding energy coincides with the
335: experimental value. This fitting of the Brueckner orbitals can be considered
336: as a way of including other higher-order diagrams into the calculations.
337: 
338: If we use Brueckner orbitals instead of RHF orbitals to calculate
339: the PNC amplitude in Eq. \ref{eq:TDHF}, then we include all-orders
340: in $\hat{\Sigma}$ contributions to $E_{PNC}$.
341: However, the correlation potential is energy-dependent,
342: $\hat{\Sigma}=\hat{\Sigma}(\epsilon)$.
343: So, in first order, we should consider the proper energy dependence.
344: The first-order in $\hat{\Sigma}$ corrections to $E_{PNC}$ are
345: presented diagrammatically in Fig. \ref{fig:pncdom}.
346: We can write these as
347: \begin{equation}
348: \langle \psi _{7s}|\hat{\Sigma}(\epsilon _{7s})|\delta X_{6s}\rangle
349: +\langle \delta\psi _{7s}|\hat{\Sigma}(\epsilon _{7s})|X_{6s}\rangle
350: +\langle \delta Y_{7s}|\hat{\Sigma}(\epsilon _{6s})|\psi_{6s}\rangle
351: +\langle Y_{7s}|\hat{\Sigma}(\epsilon _{6s})|\delta \psi_{6s}\rangle \ .
352: \end{equation}
353: The non-linear in $\hat{\Sigma}$ contribution can be found by subtracting
354: from the all-orders result the first-order value found in the same method.
355: 
356: The correlation corrections to $E_{PNC}$ we have considered so far
357: are usually called ``Brueckner-type'' corrections.
358: (In this case the external field interacts with the external electron
359: lines.) There are also contributions to $E_{PNC}$ in which the
360: external field acts inside the correlation potential
361: (see Fig. \ref{fig:pncint}).
362: Those diagrams in which the E1 interaction occurs in the internal
363: lines are known as ``structural radiation'',
364: while those in which the weak interaction occurs in the internal lines are
365: known as the ``weak correlation potential''.
366: There is another second-order correction to the amplitudes which arises
367: from the normalization of states \cite{dzuba87}.
368: The structural radiation, weak correlation potential, and
369: normalization contributions are suppressed by the small parameter
370: $E_{\rm ext}/E_{\rm core}\sim 1/10$, where  $E_{\rm ext}$ and
371: $E_{\rm int}$ are excitation energies of the external and core electrons,
372: respectively.
373: 
374: In this work we also calculate the contribution of the Breit
375: interaction.
376: We do not discuss the method of calculation here,
377: but refer the interested reader to the works \cite{dzuba01,dzuba}.
378: 
379: We postpone the discussion of our method for the calculation of radiative
380: corrections to Section \ref{sec:rad}.
381: 
382: %************************************************************
383: \section{$6s-7s$ PNC amplitude}
384: \label{sec:pnc}
385: 
386: The results of our calculation for the  $6s-7s$ PNC amplitude
387: are presented in Table~\ref{tab:pnci}.
388: Notice that the time-dependent Hartree-Fock value gives a
389: contribution to the total amplitude of about $98\%$.
390: The point is that there is a strong cancellation of the
391: correlation corrections to the PNC amplitude.
392: The stability of the PNC amplitude compared to other
393: quantities in which the correlation corrections are large
394: will be discussed in more detail in Section \ref{sec:accuracy}.
395: Notice that the values are in agreement with our 1989 result
396: $0.908\times 10^{-11}iea_{B}(-Q_{W}/N)$
397: (see ``Subtotal'' of Table \ref{tab:pnci} for the current calculation).
398: The higher numerical accuracy of the current work has therefore
399: not changed the previous result.
400: 
401: The mixed-states approach has also been performed in
402: \cite{blundell90} and \cite{kozlov01} to determine the PNC amplitude
403: in cesium.
404: However, in these works the screening of the
405: electron-electron interaction was included in a simplified way.
406: In \cite{blundell90} empirical screening factors were placed before
407: the second-order correlation corrections $\hat{\Sigma}^{(2)}$ to fit the
408: experimental values of energies.
409: Kozlov and Porsev introduced screening factors based on average screening
410: factors calculated for the Coulomb integrals between valence electron
411: states.
412: The results obtained by these groups
413: (without the Breit interaction, i.e., corresponding to the
414: Subtotal of Table \ref{tab:pnci}),
415: $0.904$ \cite{blundell90} and $0.905$ \cite{kozlov01},
416: are close to ours.
417: To be sure that we understand this difference, we performed a
418: pure second-order (i.e., using $\hat{\Sigma}^{(2)}$)
419: calculation and fitted the energies (as was done
420: in \cite{blundell90}) and reproduced their result, $0.905$.
421: 
422: In the work \cite{blundell90} a calculation using the
423: sum-over-states method was also performed.
424: In the sum-over-states approach the $6s-7s$ PNC amplitude
425: is expressed in the form
426: \begin{equation} \label{sum}
427: E_{PNC}=\sum _{n} \Big(
428: \frac{\langle 7s |\hat{H}_{W}|np\rangle \langle np|\hat{H}_{E1}|6s \rangle}
429: {E_{7s}-E_{np}} +
430: \frac{\langle 7s |\hat{H}_{E1}|np\rangle \langle np|\hat{H}_{W}|6s \rangle}
431: {E_{6s}-E_{np}} \Big) \ .
432: \end{equation}
433: The authors of reference \cite{blundell90} include single, double, and
434: selected triple excitations into their wavefunctions.
435: Note, however, that even if wavefunctions of $6s$, $7s$, and intermediate
436: $np$ states are calculated exactly
437: (i.e., with all configuration mixing included)
438: there are still some missed contributions in this approach.
439: Consider, e.g., the intermediate state $6p\equiv 5p^{6}6p$.
440: It contains an admixture of states $5p^{5}ns6d$:
441: $\tilde{6p}=5p^{6}6p + \alpha 5p^{5}ns6d+...$.
442: This mixed state is included into the  sum (\ref{sum}).
443: However, the sum (\ref{sum}) must include all many-body states
444: of opposite parity.
445: This means that the state $\tilde{5p^{5}ns6d}=
446: 5p^{5}ns6d- \alpha 5p^{6}6p+...$ should also be included into
447: the sum. Such contributions to $E_{PNC}$ have never
448: been estimated directly within the sum-over-states approach.
449: However, they are included into the mixed-states calculation.
450: The result of the sum-over-states approach, 0.909,
451: is very close to the result of the mixed-states aproach, 0.908.
452: It is important to note that the omitted higher-order many-body corrections
453: are different in these two methods.
454: This may be considered as an argument that the omitted many-body corrections
455: in both calculations are small.
456: Of course, here we assume that the omitted many-body corrections to both
457: values (which, in principle, are completely different) do not
458: ``conspire'' to give exactly the same magnitude.
459: 
460: Therefore we will take $0.908$ for the value of $E_{PNC}$
461: (Subtotal of Table \ref{tab:pnci}) as this corresponds
462: to the most complete mixed-states calculation
463: and is in agreement with the sum-over-states calculation of
464: reference \cite{blundell90}.
465: 
466: With Breit, our result becomes $0.903\times 10^{-11}iea_{B}(-Q_{W}/N)$.
467: This correction is in agreement with
468: \cite{derevianko00,dzuba01,kozlov01}.
469: 
470: We use the two-parameter Fermi model for the proton and neutron distributions:
471: \begin{equation}
472: \rho (r)=\rho _{0}\Big[ 1+\exp [(r-c)/a] \Big] ^{-1}\ ,
473: \end{equation}
474: where $t=a(4\ln 3)$ is the skin-thickness,
475: $c$ is the half-density radius, and
476: $\rho _{0}$ is found from the normalization condition $\int \rho dV=1$.
477: In 1989 the thickness and half-density radius for the proton
478: distribution were taken to be $t_{p}=2.5~{\rm fm}$ and
479: $c_{p}=5.6149~{\rm fm}$ (corresponding to a root-mean-square
480: (rms) radius $\langle r_{p}^{2}\rangle ^{1/2}=4.836~{\rm fm}$).
481: In this work we have used improved parameters
482: $t_{p}=2.3~{\rm fm}$ and $c_{p}=5.6710~{\rm fm}$
483: ($\langle r_{p}^{2}\rangle ^{1/2}=4.804~{\rm fm}$) \cite{fricke95}.
484: This changes the wavefunctions slightly,
485: leading to a very small correction to the PNC amplitude
486: of $0.08\%$ ($0.0007$).
487: (This is in agreement with a simple analytical estimate:
488: the factor accounting for the change in the electron density
489: is $\sim (4.804/4.836)^{-Z^{2}\alpha ^{2}}\sim 0.1\% \ $.)
490: In the work \cite{dzuba89} we used the proton distribution in the
491: weak interaction Hamiltonian (Eq. \ref{eq:h_w}).
492: In the current work we have found the small correction to
493: $E_{PNC}$ which arises from taking the (poorly understood)
494: neutron density in Eq. \ref{eq:h_w}.
495: We use the result of Ref. \cite{r_{np}} for the difference
496: $\Delta r_{np}=0.13(4)~{\rm fm}$
497: in the root-mean-square radii of the neutrons
498: $\langle r_{n}^{2}\rangle ^{1/2}$ and protons
499: $\langle r_{p}^{2}\rangle ^{1/2}$.
500: We have considered three cases which correspond to the same value of
501: $\langle r_{n}^{2}\rangle$: (i) $c_{n}=c_{p}$, $a_{n}>a_{p}$;
502: (ii) $c_{n}>c_{p}$, $a_{n}>a_{p}$; and (iii) $c_{n}>c_{p}$, $a_{n}=a_{p}$
503: (using the relation $\langle r_{n}^{2}\rangle \approx \frac{3}{5}c_{n}^{2}
504: +\frac{7}{5}\pi ^{2}a_{n}^{2}$).
505: We have found that $E_{PNC}$ shifts from $-0.18\%$ to $-0.21\%$
506: when moving from the extreme $c_{n}=c_{p}$ to the extreme $a_{n}=a_{p}$.
507: Therefore, $E_{PNC}$ changes by about $-0.2\%$ ($-0.0018$)
508: due to consideration of the neutron distribution.
509: This is in agreement with Derevianko's estimate,
510: $-0.19(8)\%$ \cite{derevianko}.
511: 
512: In the next section we discuss the radiative corrections to
513: $E_{PNC}$. Our estimate of these corrections does not shift
514: the PNC amplitude.
515: 
516: Therefore, we have
517: \begin{equation}
518: E_{PNC}=0.902 \times 10^{-11}iea_{B}(-Q_{W}/N)
519: \end{equation}
520: as our central point for the PNC amplitude.
521: The error will be estimated in the following sections.
522: 
523: \section{QED-type radiative corrections to energy levels,
524: wavefunctions, and the PNC amplitude}
525: \label{sec:rad}
526: 
527: The radiative corrections to the weak charge $Q_W$ have been
528: calculated for the free electron. However, an electron in a heavy
529: atom is bound, and this produces additional radiative
530: corrections proportional to $\alpha (Z \alpha)^n$,
531: $n=1,2,...$. Recently such corrections were considered
532: by Milstein and Sushkov \cite{milstein}. They found that
533: the most important are corrections enhanced by
534: the large parameter $\ln(\lambda/R)$, where $\lambda=\hbar/mc$
535: is the electron Compton wavelength and $R$ is the nuclear
536: radius. This type of correction arises from the radiative
537: corrections to the electron wavefunction near the nucleus.
538: In this region the $s$-wave and $p_{1/2}$-wave (lower Dirac component)
539: electron densities are singular, $|\psi(r)|^2 \sim r^{-Z^2\alpha^2}$.
540: The radiative corrections modify the potential at small distances
541: $r<\lambda$, ${\tilde V }(r)= -Z \alpha (1+\delta)/r$.
542: Correspondingly, the electron wavefunctions change,
543: $|\psi(r)|^2 \sim r^{-Z^2\alpha^2 (1 +\delta)^2}$ for $r<\lambda$.
544: This gives the radiative correction factor for the electron
545: density inside the nucleus,
546: \begin{equation}
547: \label{psirad}
548: \frac{|\psi(R)|^2}{|\psi(\lambda)|^2} \sim \Big( \frac{\lambda}{R}\Big)
549: ^{Z^2\alpha^2 2\delta}=\exp \Big( 2\delta Z^2\alpha^2 \ln(\lambda /R) \Big)
550: \ .
551: \end{equation}
552: For the Uehling (vacuum polarization) potential
553: $\delta \sim \alpha \ln(\lambda /r)$ \cite{berestetskii}.
554: This gives an additional power of the large  parameter $\ln(\lambda / R)$.
555: This leads Milstein and Sushkov \cite{milstein} to conclude that the
556: Uehling potential gives a dominating radiative correction
557: to $E_{PNC}$, $\sim  Z^2\alpha^3 \ln^2(\lambda/ R)$.
558: Numerical calculations of the Uehling potential contribution
559: have been performed in \cite{johnson01} and in the present work.
560: This radiative correction increases $E_{PNC}$ by 0.4\%.
561: 
562: Milstein and Sushkov \cite{milstein} demonstrated that there are no other
563: radiative corrections which are enhanced by  $\ln^2(\lambda /R)$.
564: However, any correction to the potential with nonzero
565: $\delta(R) \sim \alpha$ gives a correction to the electron density
566: $\sim  Z^2\alpha^3 \ln (\lambda /R)$. We demonstrate below
567: that such corrections are also important and give a contribution
568: of opposite sign to that of the Uehling potential.
569: 
570: Let us start our discussion from the radiative corrections to energy
571: levels (the Lamb shift).
572: The calculation of the shift can be divided into two parts:
573: one in which the electron interaction with virtual photons of
574: high-frequency are considered, and one in which
575: virtual photons of low-frequency are considered.
576: 
577: In the high-frequency case the external field
578: (the strong nuclear Coulomb field) need only be included
579: to first order.
580: In this case the contributions to the Lamb shift arise
581: from the diagrams presented in Fig. \ref{fig:rad}.
582: The contribution of the Uehling potential (Fig. \ref{fig:rad}(a))
583: to the Lamb shift is very small.
584: The main contribution comes from the vertex correction
585: (Fig. \ref{fig:rad}(b)).
586: (In the case of a free electron the vertex diagrams give the electric
587: $f(q^{2})$ and magnetic $g(q^{2})$ formfactors.)
588: The perturbation theory expression for $f(q^2)$ contains an
589: infra-red divergence and requires a low-frequency
590: cut-off parameter $\kappa$ - see, e.g., \cite{berestetskii}.
591: Assuming $q^2 \ll m^2 c^2$, the high-frequency contribution to the
592: Lamb shift can be presented as a potential given by the following
593: expression \cite{berestetskii}
594: \begin{eqnarray}
595: \delta \Phi ({\bf r})
596: &=&\Big[ \delta \Phi _{f}+\delta \Phi _{U}\Big] +\delta \Phi _{g}\nonumber \\
597: &=&\frac{\alpha \hbar ^{2}}{3\pi m^{2}c^{2}}\Big(
598: \ln \frac{m}{2\kappa} +\frac{11}{24}-\frac{1}{5}\Big)
599: \Delta \Phi ({\bf r})-
600: i\frac{\alpha \hbar}{4\pi mc}\mbox{\boldmath$\gamma$}\cdot
601: \mbox{\boldmath$\nabla$}\Phi ({\bf r}) \ .
602: \end{eqnarray}
603: For the Coulomb potential, $\Delta \Phi =-4\pi Ze\delta ({\bf r})$.
604: Here the last long-range term ($\delta \Phi _{g}$) comes from
605: the anomalous electron magnetic moment ($g(0)$);
606: the infra-red cut-off parameter $\kappa$ appears from the electric
607: formfactor $f(q^{2})$.
608: This term with the large $\ln \frac{m}{2\kappa}$ gives the dominant
609: contribution to the Lamb shift of $s$-levels.
610: The infra-red divergence for $\kappa \rightarrow 0$ indicates the
611: importance of the low-frequency contribution for this term.
612: 
613: If we go beyond the approximation
614: $q^{2}<<m^{2}c^{2}$, the term $\delta \Phi _{f}$ should be
615: associated with a non-local self-energy operator
616: ${\hat \Sigma}_{\rm rad} ({\bf r},{\bf r}',E)$ with typical values
617: $|{\bf r}-{\bf r}'|\lesssim \frac{\hbar}{mc}$
618: and $r\sim r' \lesssim \frac{\hbar}{mc}$.
619: We need this operator in a simple limit, $E<<mc^{2}$.
620: In this case we can approximate it by a two-parametric $(A,b)$ potential
621: \begin{equation}
622: \delta \Phi _{f}=-A\frac{\alpha}{\pi} {\rm e}^{-b\frac{mc}{\hbar}r}\Phi \ .
623: \end{equation}
624: The parameters $A$ and $b$ in $\delta \Phi _{f}$ can be found from the
625: fit of the Lamb-shift of the high Coulomb levels $3s$, $4s$, $5s$ and
626: $3p$, $4p$ and $5p$ (in one-electron ions) which were calculated as
627: a function of the nuclear charge $Z$ in Refs. \cite{mohr}.
628: We have checked that $A=1.25$ and $b=1$ fit all these
629: Lamb shifts quite accurately (we have found slightly different potential
630: strengths for $s$- and $p$-waves, $A_s=1.17$ and $A_p=1.33$).
631: The anomalous magnetic moment contribution is
632: \begin{equation}
633: \delta \Phi _{g}=-i\frac{\alpha \hbar}{4\pi mc}\mbox{\boldmath$\gamma$}
634: \cdot \mbox{\boldmath$\nabla$}\Phi \ .
635: \end{equation}
636: The Uehling potential for a finite nucleus is given by \cite{fullerton76}
637: (in atomic units ($\hbar =m=e=1$, $\alpha =1/c$))
638: \begin{equation}
639: \delta \Phi _{U}=-\frac{2\alpha ^{2}}{3r}\int _{0}^{\infty}dx~
640: x\rho(x)\int _{1}^{\infty}dt~\sqrt{t^{2}-1}
641: \Big(
642: \frac{1}{t^{3}}+\frac{1}{2t^{5}}\Big)
643: \Big(
644: {\rm e}^{-2t|r-x|/\alpha} -{\rm e}^{-2t(r+x)/\alpha} \Big)    \ ,
645: \end{equation}
646: where $\rho (x)$ is the nuclear charge density.
647: It is more convenient to use a simpler formula for
648: $\delta \Phi _{U}$ for $r\geq R$, $R$ is the nuclear radius,
649: \begin{eqnarray}
650: &&\delta \Phi _{U}(r)= \Phi (r)\frac{\alpha ^{4}}{8\pi R^{3}}
651: \int _{1}^{\infty}dt~ \sqrt{t^{2}-1}
652: \Big(
653: \frac{1}{t^{5}}+\frac{1}{2t^{7}}\Big)
654: {\rm e}^\frac{-2tr}{\alpha}I(x)  \ , \\
655: &&I(x)=-{\rm e}^{x}+{\rm e}^{-x}+x{\rm e}^{x} +x{\rm e}^{-x} \ ,
656: \qquad x=2tR/\alpha \ ,
657: \end{eqnarray}
658: and take $\delta \Phi _{U}(r<R)=\delta \Phi _{U}(r=R)$.
659: There is practically no loss of numerical accuracy in this
660: approximation since a typical scale for the variation of
661: $\delta \Phi _{U}(r)$ is given by the electron compton length
662: $\frac{\hbar}{mc}>>R$.
663: The radiative corrections for Cs energy levels are presented
664: in Table \ref{tab:radenergies}.
665: 
666: Note that for the most important term $\delta \Phi _{f}$ we do not
667: use the assumption $Z\alpha <<1$ since we fitted the exact results
668: for the single-electron ions.
669: The contribution of the Uehling potential $\delta \Phi _{U}$ to the Lamb
670: shift is always small.
671: Also, Milstein and Strakhovenko have shown in Ref. \cite{milstein83}
672: that higher $Z\alpha$ corrections are numerically not important
673: for this potential.
674: The potential $\delta \Phi _{g}$ due to the magnetic formfactor
675: is a long-range one.
676: This also hints that there are no large higher $Z\alpha$
677: corrections here.
678: 
679: We can use $\delta \Phi$ to estimate the contribution of
680: QED-type radiative corrections to the electron wavefunction
681: and  PNC amplitude $E_{PNC}$.
682: Note that it is not enough to calculate the radiative corrections to
683: the matrix element of the weak interaction
684: $\langle n'p_{1/2}|\hat{H}_{W}|ns\rangle$.
685: Corrections to the energy intervals like $6s-6p$  are also important
686: since these intervals are small at the scale of the atomic unit
687: ($ \sim 1/20$) and sensitive to perturbations.
688: The change in the energies also influences the large-distance behavior
689: of the electron wavefunctions which determine the usual
690: E1 amplitudes in the sum-over-states approach.
691: Thus, the simplest way to proceed is to include $\delta \Phi$ into the
692: Dirac-Hartree-Fock equations and then perform all calculations.
693: 
694: The results are the following:
695: the Uehling potential $\delta \Phi _{U}$ increases $E_{PNC}$ by
696: $0.41\%$ (in agreement with \cite{johnson01});
697: the contribution of $\delta \Phi _{g}$ is small, $-0.03 \%$
698: (due to cancellation of the contributions of the corrections
699: to the $s$-wave and $p$-wave);
700: and the contribution of $\delta \Phi _{f}$ is $-0.65\%$.
701: The total result of $\delta \Phi$ is $-0.27\%$
702: (and opposite to that of the Uehling potential).
703: 
704: Note that the replacement of
705: ${\hat \Sigma}_{\rm rad} ({\bf r},{\bf r}',E=0)$ by a
706: local parametric potential ($\delta \Phi _{f}$) may be quite a crude
707: approximation. Therefore, we have tested the sensitivity of
708: the result to variation of the parameter $b$
709: (parameter $A$ is a function of $b$, it is found from the fit
710: of the Coulomb energy levels).
711: The increase of the radius of the potential $\delta \Phi _{f}$ two times
712: ($b=0.5$) increases $E_{PNC}$ by $0.3\%$
713: (the total contribution of $\delta \Phi$ is $0.04\%$).
714: 
715: We do not discuss here the $Z\alpha$ correction to the
716: $Z$-boson exchange vertex. Milstein and Sushkov \cite{milstein} have
717: shown that this correction does not contain the large
718: parameter $\ln(\lambda /R)$ if the calculations
719: are performed in the Landau gauge and they concluded
720: that this correction should be small.
721: 
722: Here we should note that the radiative corrections to
723: the wavefunctions (as well as the wavefunctions themselves)
724: are not gauge invariant.
725: However, the change of the electron density inside the nucleus is
726: a gauge invariant, observable phenomenon.
727: This makes the consideration presented above meaningful.
728: 
729: Of course, our estimate of QED-type radiative corrections to
730: $E_{PNC}$ should not be considered as an accurate calculation.
731: The aim is to show that the consideration of just the single
732: Uehling potential contribution to the radiative corrections
733: can give the wrong impression.
734: Indeed, this contribution to $E_{PNC}$ is enhanced by the large
735: parameter $\ln^2(\lambda /R)$.
736: However, this only makes it comparable to other contributions which
737: in the case of energy levels were an order of magnitude
738: larger than that of the Uehling potential.
739: A conservative estimate of the radiative correction contribution
740: can be presented as $0.0 \pm 0.4 \%$.
741: This range covers our two values (-0.27\% and 0.04\%) as well as
742: the value 0.4\% obtained in \cite{milstein,johnson01}.
743: 
744: %------------------------------------------------------------
745: 
746: \section{Estimate of accuracy of PNC amplitude}
747: \label{sec:accuracy}
748: 
749: We have estimated the error of the PNC amplitude in a number
750: of different ways. There are two main methods:
751: (i) root-mean-square (rms) deviation of calculated energy intervals,
752: E1 amplitudes, and hyperfine structure (hfs) constants
753: from the accurate experimental values;
754: (ii) influence of fitting of energies and hyperfine structure
755: constants on the PNC amplitude.
756: 
757: \subsection{Root-mean-square deviation}
758: 
759: Remember that the PNC amplitude can be expressed as a sum over
760: intermediate states (see formula \ref{sum}).
761: Each term in the sum is a product of E1 transition amplitudes,
762: weak matrix elements, and energy denominators.
763: There are three dominating contributions to the $6s-7s$ PNC amplitude
764: in Cs \cite{blundell90}:
765: \begin{eqnarray}
766: E_{PNC}&=&
767: \label{sumcs}
768: \frac{\langle 7s|\hat{H}_{E1}|6p\rangle \langle 6p|\hat{H}_{W}|6s\rangle}
769: {E_{6s}-E_{6p}} +
770: \frac{\langle 7s|\hat{H}_{W}|6p\rangle \langle 6p|\hat{H}_{E1}|6s\rangle}
771: {E_{7s}-E_{6p}}+
772: \frac{\langle 7s|\hat{H}_{E1}|7p\rangle \langle 7p|\hat{H}_{W}|6s\rangle}
773: {E_{6s}-E_{7p}}+...
774: \nonumber \\
775: &=& -1.908+1.493+1.352+...=0.937 +... \ .
776: \end{eqnarray}
777: While we do not use the sum-over-states approach in our calculation of
778: the PNC amplitude, it is instructive to analyze the accuracy of
779: the E1 transition amplitudes, weak matrix elements, and
780: energy intervals which contribute to Eq. \ref{sumcs} as
781: they have been calculated using the same method as that used to
782: calculate $E_{PNC}$.
783: 
784: Let us begin with the energy intervals.
785: The calculated ionization energies are presented in Table \ref{tab:energies}.
786: The Hartree-Fock values deviate from experiment by $10\%$.
787: Including the second-order correlation corrections ${\hat \Sigma} ^{(2)}$
788: reduces the error to $\sim 1\%$.
789: When screening and the hole-particle interaction are included into
790: ${\hat \Sigma} ^{(2)}$ in all orders,
791: the energies improve, $\sim 0.1\%$.
792: The percentage deviations from experiment of the energy intervals of interest
793: are: $E_{6s}-E_{6p}$, $-0.3$; $E_{7s}-E_{6p}$, $0.8$;
794: and $E_{6s}-E_{7p}$, $-0.01$. The rms error is $0.5\%$.
795: We can in fact reproduce energy intervals exactly by placing coefficients
796: before the correlation potential. Because this fitting of the
797: energies appears to improve the wavefunctions (e.g., electromagnetic
798: amplitudes and hyperfine structure constants improve)
799: we will use this procedure in the following estimates.
800: 
801: The relevant radial integrals (E1 transition amplitudes) are presented in
802: Table \ref{tab:e1i}.
803: These were calculated with the energy-fitted ``bare'' correlation
804: potential $\hat{\Sigma}^{(2)}$ and the (unfitted and fitted)
805: ``dressed'' potential $\hat{\Sigma}$.
806: Structural radiation and normalization contributions were also
807: included.
808: In Table \ref{tab:e1ii} the percentage deviations of
809: the calculated values from experiment are listed.
810: Without energy fitting, the rms error is $0.3\%$.
811: Fitting the energy improves the accuracy, $\sim 0.1\%$.
812: 
813: We cannot directly compare weak matrix elements with experiment.
814: Like the weak matrix elements, hyperfine structure is determined
815: by the electron wavefunctions in the vicinity of the nucleus,
816: and this is known very accurately.
817: The hyperfine structure constants calculated in different approximations
818: are presented in Table \ref{tab:hfsi}.
819: Corrections due to the Breit interaction, structural radiation,  and
820: normalization are included.
821: The percentage deviations from experiment are shown in
822: Table \ref{tab:hfsii}.
823: The rms deviation of the calculated hfs values from experiment
824: using unfitted ${\hat \Sigma}$ is $1.1\%$.
825: With fitting, the rms error in the pure second-order approximation is $0.3\%$;
826: with higher orders we get $0.6\%$.
827: We are, however, trying to estimate the accuracy of the $s-p$ weak
828: matrix elements. It makes more sense for us to use the square-root
829: formula, $\sqrt{{\rm hfs}(s){\rm hfs}(p)}$. The errors are presented in Table
830: \ref{tab:hfsii}.
831: Notice that by using this approach the error is much smaller.
832: Without energy fitting, the rms error is $0.4\%$.
833: With fitting, the rms error in the second-order calculation
834: ($\hat{\Sigma} ^{(2)}$) and full calculation
835: ($\hat{\Sigma}$) is $0.2\%$.
836: 
837: From this section we can conclude that the rms error for the relevant
838: parameters is somewhere in the range $0.2-0.5\%$.
839: 
840: Note that from this analysis the error for the sum-over-states
841: calculation of $E_{PNC}$ would be larger than this, as
842: the errors for the energies, hfs constants, and E1 amplitudes
843: contribute to each of the three terms in Eq. \ref{sumcs}.
844: However, in the mixed-states approach, the errors do not add in this
845: way.
846: We get a better indication of the error of our calculation
847: of $E_{PNC}$ in the next section.
848: 
849: 
850: \subsection{Influence of fitting on the PNC amplitude}
851: 
852: In the section above we presented calculations in three different
853: approximations:
854: with unfitted $\hat{\Sigma}$,
855: and with $\hat{\Sigma}^{(2)}$ and $\hat{\Sigma}$ fitted with
856: coefficients to reproduce experimental ionization energies.
857: The errors in these approximations are of different magnitudes and signs.
858: We now calculate the PNC amplitude using these three approximations.
859: The spread of the results can be used to estimate the error.
860: 
861: The results are listed in Table \ref{tab:pncii}.
862: It can be seen that the PNC amplitude is very stable.
863: 
864: The PNC amplitude is much more stable than hyperfine structure.
865: This can be explained by the much smaller correlation corrections
866: to PNC (compare Table \ref{tab:pnci} with Table \ref{tab:hfsi}).
867: The stability of $E_{PNC}$ may be compared to the stability of the
868: usual electromagnetic amplitudes where the error is very small
869: (even without fitting).
870: 
871: We have also considered the fitting of hyperfine structure using
872: different coefficients before each $\hat{\Sigma}$. Using the resulting
873: wavefunctions, the PNC amplitude increased by $0.5\%$.
874: This is the maximum deviation we have obtained. We will therefore
875: use this as the estimate for the accuracy of the $E_{PNC}$ calculation.
876: 
877: %------------------------------------------------------------
878: 
879: \section{Conclusion}
880: 
881: We have obtained the result
882: \begin{equation}
883: E_{PNC}=0.902(6) \times 10^{-11}iea_{B}(-Q_{W}/N)
884: \end{equation}
885: for our calculation of the $6s-7s$ PNC amplitude in Cs.
886: This is in agreement with other PNC calculations,
887: however we would like to emphasize that our calculation is
888: the most complete.
889: The most precise measurement of the $6s-7s$ PNC amplitude in Cs
890: is \cite{wood97}
891: \begin{equation}
892: -\frac{{\rm Im} (E_{PNC})}{\beta}=1.5939(56)\frac{\rm mV}{\rm cm} \ ,
893: \end{equation}
894: where $\beta$ is the vector transition polarizability.
895: For $\beta$ we use the value
896: \begin{equation}
897: \beta =27.00(6) a_{B}^{3}
898: \end{equation}
899: which is the average value of the most accurate
900: results \cite{dzuba00,wieman99} and \cite{dzuba97,wood97}.
901: Using the conversion $|e|/a_{B}^{2}=5.1422\times 10^{12}{\rm mV}/{\rm cm}$,
902: we therefore obtain for the weak charge of the Cs nucleus:
903: \begin{equation}
904: Q_{W}=-72.39(29)_{\rm exp}(51)_{\rm theory} \ ,
905: \end{equation}
906: where the experimental error is obtained by adding in quadrature the
907: error for $\beta$ and the error for ${\rm Im}(E_{PNC})/\beta$.
908: This result deviates by
909: $\Big( 1.0 \pm 0.4({\rm exp})\pm 0.7({\rm theory})\Big) \%$ from
910: the Standard Model value $Q_{W}= -73.09(3)$ \cite{groom00}.
911: This does not represent a significant deviation.
912: 
913: \acknowledgments
914: 
915: We are grateful to A. Milstein, O. Sushkov, and M. Kuchiev for
916: useful discussions.
917: This work was supported by the Australian Research Council.
918: 
919: %************************************************************
920: \begin{thebibliography}{20}
921: 
922: \bibitem{dzuba89}
923: 
924: V.A. Dzuba, V.V. Flambaum, and O.P. Sushkov,
925: Phys. Lett. A {\bf 141}, 147 (1989).
926: 
927: \bibitem{blundell90}
928: 
929: S.A. Blundell, W.R. Johnson, and J. Sapirstein,
930: Phys. Rev. Lett. {\bf 65}, 1411 (1990);
931: S.A. Blundell, J. Sapirstein, and W.R. Johnson,
932: Phys. Rev. D {\bf 45}, 1602 (1992).
933: 
934: \bibitem{wood97}
935: 
936: C.S. Wood {\it et al.}, Science {\bf 275}, 1759 (1997).
937: 
938: \bibitem{wieman99}
939: 
940: S.C. Bennett and C.E. Wieman, Phys. Rev. Lett. {\bf 82},
941: 2484 (1999); {\bf 83}, 889 (1999).
942: 
943: \bibitem{derevianko00}
944: 
945: A. Derevianko, Phys. Rev. Lett. {\bf 85}, 1618 (2000).
946: 
947: \bibitem{sushkov01}
948: 
949: O.P. Sushkov, Phys. Rev. A {\bf 63}, 042504 (2001).
950: 
951: \bibitem{dzuba01}
952: 
953: V.A. Dzuba, C. Harabati, W.R. Johnson, and M.S. Safronova,
954: Phys. Rev. A {\bf 63}, 044103 (2001).
955: 
956: \bibitem{kozlov01}
957: 
958: M.G. Kozlov, S.G. Porsev, and I.I. Tupitsyn,
959: Phys. Rev. Lett. {\bf 86}, 3260 (2001).
960: 
961: \bibitem{marciano}
962: 
963: W.J. Marciano and A. Sirlin, Phys. Rev. D {\bf 27}, 552 (1983);
964: W.J. Marciano and J.L. Rosner, Phys. Rev. Lett. {\bf 65}, 2963 (1990).
965: 
966: \bibitem{lynnbed}
967: 
968: B.W. Lynn and P.G.H. Sandars, J. Phys. B. {\bf 27}, 1469 (1994);
969: I. Bednyakov, L. Labzowsky, G. Plunien, G. Soff, and V. Karasiev,
970: Phys. Rev. A {\bf 61}, 012103 (1999).
971: 
972: \bibitem{milstein}
973: 
974: A.I. Milstein and O.P. Sushkov, e-print hep-ph/0109257.
975: 
976: \bibitem{johnson01}
977: 
978: W.R. Johnson, I. Bednyakov, and G. Soff, submitted to Phys. Rev. Lett.;
979: e-print hep-ph/0110262.
980: 
981: \bibitem{derevianko}
982: 
983: A. Derevianko, submitted to Phys. Rev. A; e-print physics/0108033.
984: 
985: \bibitem{dzuba00}
986: 
987: V.A. Dzuba and V.V. Flambaum,
988: Phys. Rev. A {\bf 62}, 052101 (2000).
989: 
990: \bibitem{dzuba97}
991: 
992: V.A. Dzuba, V.V. Flambaum, and O.P. Sushkov,
993: Phys. Rev. A {\bf 56}, R4357 (1997).
994: 
995: \bibitem{groom00}
996: 
997: D.E. Groom {\it et al.},
998: Euro. Phys. J. C {\bf 15}, 1 (2000).
999: 
1000: \bibitem{dzuba89energy}
1001: 
1002: V.A. Dzuba, V.V. Flambaum, and O.P. Sushkov,
1003: Phys. Lett. A {\bf 140}, 493 (1989).
1004: 
1005: \bibitem{dzuba89e1hfs}
1006: 
1007: V.A. Dzuba, V.V. Flambaum, A.Ya. Kraftmakher, and O.P. Sushkov,
1008: Phys. Lett. A {\bf 142}, 373 (1989).
1009: 
1010: \bibitem{dzuba87}
1011: 
1012: V.A. Dzuba, V.V. Flambaum, P.G. Silvestrov, and O.P. Sushkov,
1013: J. Phys. B {\bf 20}, 1399 (1987).
1014: 
1015: \bibitem{dzuba}
1016: 
1017: V.A. Dzuba, in preparation.
1018: 
1019: %-------------
1020: %proton distrn
1021: 
1022: \bibitem{fricke95}
1023: 
1024: G. Fricke {\it et al.},
1025: At. Data and Nucl. Data Tables {\bf 60}, 177 (1995).
1026: 
1027: %-------------
1028: %neutron distrn
1029: 
1030: \bibitem{r_{np}}
1031: 
1032: A. Trzci\'{n}ska {\it et al.},
1033: Phys. Rev. Lett. {\bf 87}, 082501 (2001).
1034: 
1035: \bibitem{berestetskii}
1036: 
1037: V.B. Berestetskii, E.M. Lifshitz, and L.P. Pitaevskii,
1038: {\it Relativistic Quantum Theory}
1039: (Pergamon Press, Oxford, 1982).
1040: 
1041: \bibitem{mohr}
1042: 
1043: P.J. Mohr and Y.-K. Kim, Phys. Rev. A {\bf 45}, 2727 (1992);
1044: P.J. Mohr, Phys. Rev. A {\bf 46}, 4421 (1992).
1045: 
1046: \bibitem{fullerton76}
1047: 
1048: L.W. Fullerton and G.A. Rinker, Jr., Phys. Rev. A {\bf 13},
1049: 1283 (1976).
1050: 
1051: \bibitem{milstein83}
1052: 
1053: A.I. Milstein and V.M. Strakhovenko,
1054: ZhETF {\bf 84}, 1247 (1983).
1055: 
1056: %------------------
1057: % energies
1058: 
1059: \bibitem{moore}
1060: 
1061: C.E. Moore, Natl. Stand. Ref. Data Ser.
1062: (U.S., Natl. Bur. Stand.), {\bf 3} (1971).
1063: 
1064: %------------------
1065: % radial integrals
1066: 
1067: \bibitem{rafac99}
1068: 
1069: R.J. Rafac, C.E. Tanner, A.E. Livingston, and H.G. Berry,
1070: Phys. Rev. A {\bf 60}, 3648 (1999).
1071: 
1072: \bibitem{bouchiat84}
1073: 
1074: M.-A. Bouchiat, J. Guena, and L. Pottier,
1075: J. Phys. (France) Lett. {\bf 45}, L523 (1984).
1076: 
1077: \bibitem{bennett99}
1078: 
1079: S.C. Bennett, J.L. Roberts, and C.E. Wieman,
1080: Phys. Rev. A {\bf 59}, R19 (1999).
1081: 
1082: %------------------
1083: % hfs
1084: 
1085: \bibitem{cshfs}
1086: 
1087: E. Arimondo, M. Inguscio, and P. Violino,
1088: Rev. Mod. Phys. {\bf 49}, 31 (1977).
1089: 
1090: \bibitem{gilbert83}
1091: 
1092: S.L. Gilbert, R.N. Watts, and C.E. Wieman,
1093: Phys. Rev. A {\bf 27}, 581 (1983).
1094: 
1095: \bibitem{rafac97}
1096: 
1097: R.J. Rafac and C.E. Tanner, Phys. Rev. A {\bf 56}, 1027 (1997).
1098: 
1099: \end{thebibliography}
1100: 
1101: %************************************************************
1102: \begin{table}
1103: \caption{Contributions to the $6s-7s$ $E_{PNC}$ amplitude
1104: for Cs in units $10^{-11}iea_{B}(-Q_{W}/N)$.
1105: ($\hat{\Sigma}$ corresponds to the (unfitted) ``dressed''
1106: self-energy operator.) }
1107: \label{tab:pnci}
1108: \begin{tabular}{ld}
1109: TDHF & 0.8885 \\
1110: $\langle \psi _{7s}|\hat{\Sigma}|\delta X_{6s}\rangle$ & 0.0705 \\
1111: $\langle \delta\psi _{7s}|\hat{\Sigma}|X_{6s}\rangle$ & 0.1857 \\
1112: $\langle \delta Y_{7s}|\hat{\Sigma}|\psi_{6s}\rangle$ & -0.0760 \\
1113: $\langle Y_{7s}|\hat{\Sigma}|\delta \psi_{6s}\rangle$ & -0.1420 \\
1114: Nonlinear in $\hat{\Sigma}$ correction & -0.0198 \\
1115: Weak correlation potential & 0.0038 \\
1116: Structural radiation & 0.0025 \\
1117: Normalization & -0.0049 \\
1118:  & \\
1119: Subtotal & 0.9084 \\
1120:  & \\
1121: Breit & -0.0055 \\
1122: New proton distribution & 0.0007 \\
1123: Neutron distribution & -0.0018 \\
1124: Radiative corrections & 0.0\\
1125:  & \\
1126: Total & 0.902 \\
1127: \end{tabular}
1128: \end{table}
1129: %****************************************************************
1130: \begin{table}
1131: \caption{Radiative corrections to RHF ionization energies; units $-$cm$^{-1}$.
1132: (See also Table \ref{tab:energies}.) }
1133: \label{tab:radenergies}
1134: \begin{tabular}{dddd}
1135: $6s$ & $7s$ & $6p_{1/2}$ & $7p_{1/2}$ \\
1136: \hline
1137: 
1138: -18.4 & -5.0 & 0.88 &
1139: 0.31 \\
1140: \end{tabular}
1141: \end{table}
1142: %****************************************************************
1143: \begin{table}
1144: \caption{Ionization energies for Cs in units $-$cm$^{-1}$.}
1145: \label{tab:energies}
1146: \begin{tabular}{lllll}
1147: State & RHF & $\hat{\Sigma} ^{(2)}$ & $\hat{\Sigma}$ &
1148: Experiment \tablenotemark[1]\\
1149: \hline
1150: $6s$ & 27954 & 32415 & 31420 & 31407 \\
1151: $7s$ & 12112 & 13070 & 12863 & 12871 \\
1152: $6p_{1/2}$ & 18790 & 20539 & 20276 & 20228 \\
1153: $7p_{1/2}$ & 9223 & 9731 & 9657 & 9641 \\
1154: \end{tabular}
1155: \tablenotetext[1]{Taken from \cite{moore}.}
1156: \end{table}
1157: %************************************************************
1158: \begin{table}
1159: \caption{Radial integrals of E1 transition amplitudes for Cs in
1160: different approximations. The experimental values are listed in
1161: the last column. (a.u.)}
1162: \label{tab:e1i}
1163: \begin{tabular}{ldddddd}
1164: Transition & RHF& TDHF & $\hat{\Sigma} ^{(2)}$ &
1165: $\hat{\Sigma}$ & $\hat{\Sigma}$ & Experiment \\
1166:        &      &              & with fitting &           & with fitting &
1167:  \\
1168: \hline
1169: $6s-6p$ & 6.464 & 6.093 & 5.499 & 5.510 & 5.512 &
1170: 5.497(8) \tablenotemark[1] \\
1171: $7s-6p$ & 5.405 & 5.450 & 5.198 & 5.165 & 5.201 &
1172: 5.185(27) \tablenotemark[2]\\
1173: $7s-7p$ & 13.483 & 13.376 & 12.602 & 12.641 & 12.612 &
1174: 12.625(18) \tablenotemark[3] \\
1175: \end{tabular}
1176: \tablenotetext[1]{Ref. \cite{rafac99}.}
1177: \tablenotetext[2]{Ref. \cite{bouchiat84}.}
1178: \tablenotetext[3]{Ref. \cite{bennett99}.}
1179: \end{table}
1180: %************************************************************
1181: \begin{table}
1182: \caption{Percentage deviation from experiment of calculated radial integrals
1183: in different approximations.}
1184: \label{tab:e1ii}
1185: \begin{tabular}{lddd}
1186: Transition &
1187: \multicolumn{3}{c}{Percentage deviation} \\
1188:  & $\hat{\Sigma}^{(2)}$ & $\hat{\Sigma}$ & $\hat{\Sigma}$ \\
1189:  & with fitting    &          & with fitting \\
1190: \hline
1191: $6s-6p$ & 0.04 & 0.2 & 0.3 \\
1192: $7s-6p$ & 0.3 & -0.4 & 0.3 \\
1193: $7s-7p$ & -0.2 & 0.1 & -0.1 \\
1194: \end{tabular}
1195: \end{table}
1196: %************************************************************
1197: \begin{table}
1198: \caption{Calculations of the hyperfine structure of Cs
1199: in different approximations. In the last column the experimental
1200: values are listed. Units: MHz.}
1201: \label{tab:hfsi}
1202: \begin{tabular}{ldddddd}
1203: State & RHF & TDHF & $\hat{\Sigma} ^{(2)}$ &
1204: $\hat{\Sigma}$ & $\hat{\Sigma}$ & Experiment \\
1205:        &    &     & with fitting &      & with fitting & \\
1206: \hline
1207: $6s$ & 1425.0 & 1717.5 & 2306.9 & 2287.9 & 2285.7 &
1208: 2298.2 \tablenotemark[1] \\
1209: $7s$ & 391.6 & 471.1 & 544.4 & 539.6 & 540.2 &
1210: 545.90(9) \tablenotemark[2] \\
1211: $6p_{1/2}$ & 160.9 & 200.3 & 291.5 & 296.4 & 293.3 &
1212: 291.89(8) \tablenotemark[3] \\
1213: $7p_{1/2}$ & 57.6 & 71.2 & 94.3 & 95.4 & 94.6 &
1214: 94.35 \tablenotemark[1] \\
1215: \end{tabular}
1216: \tablenotetext[1]{Ref. \cite{cshfs}.}
1217: \tablenotetext[2]{Ref. \cite{gilbert83}.}
1218: \tablenotetext[3]{Ref. \cite{rafac97}.}
1219: \end{table}
1220: %************************************************************
1221: \begin{table}
1222: \caption{Percentage deviation from experiment of calculated
1223: hyperfine structure constants in different approximations.}
1224: \label{tab:hfsii}
1225: \begin{tabular}{ldddd}
1226: State &
1227: \multicolumn{3}{c}{Percentage deviation} \\
1228:  & $\hat{\Sigma}^{(2)}$ & $\hat{\Sigma}$ & $\hat{\Sigma}$ \\
1229:  & with fitting    &          & with fitting \\
1230: \hline
1231: $6s$ & 0.4 & -0.4 & -0.5 \\
1232: $7s$ & -0.3 & -1.2 & -1.0 \\
1233: $6p_{1/2}$ & -0.1 & 1.5 & 0.5 \\
1234: $7p_{1/2}$ & 0.05 & 1.1 & 0.3 \\
1235: \end{tabular}
1236: \end{table}
1237: %****************************************************************
1238: \begin{table}
1239: \caption{Percentage deviation from experiment of calculated
1240: $\sqrt{{\rm hfs}(s){\rm hfs}(p)}$ (we will denote this by $s-p$ in
1241: the tables) in different approximations.}
1242: \label{tab:hfsiii}
1243: \begin{tabular}{lddd}
1244: $\sqrt{{\rm hfs}(s){\rm hfs}(p)}$&\multicolumn{3}{c}{Percentage deviation}\\
1245:  & $\hat{\Sigma} ^{(2)}$ & $\hat{\Sigma}$ & $\hat{\Sigma}$ \\
1246:  & with fitting    &          & with fitting \\
1247: \hline
1248: $6s-6p$ & 0.1 & 0.5 & -0.02 \\
1249: $6s-7p$ & 0.2 & -0.3 & -0.2 \\
1250: $7s-6p$ & -0.2 & 0.2 & -0.3 \\
1251: \end{tabular}
1252: \end{table}
1253: %****************************************************************
1254: \begin{table}
1255: \caption{Values for $E_{PNC}$ in different approximations;
1256: units $10^{-11}iea_{B}(-Q_{W}/N)$.}
1257: \label{tab:pncii}
1258: \begin{tabular}{cddd}
1259:  & $\hat{\Sigma}^{(2)}$ with fitting & $\hat{\Sigma}$ &
1260: $\hat{\Sigma}$ with fitting \\
1261: \hline
1262: $E_{PNC}$ & 0.898 & 0.902 & 0.901 \\
1263: \end{tabular}
1264: \end{table}
1265: %****************************************************************
1266: 
1267: \center
1268: \widetext
1269: \input psfig
1270: \psfull
1271: 
1272: \begin{figure}[b]
1273: \centerline{\psfig{file=pnc.eps,clip=}}
1274: \caption{Lowest-order diagrams for PNC.
1275: Solid line denotes the bound electron;
1276: cross is the weak interaction; and dashed line is the E1 field.}
1277: \label{fig:pnc}
1278: \end{figure}
1279: 
1280: \begin{figure}[b]
1281: \centerline{\psfig{file=corepol.eps, clip=}}
1282: \caption{Examples of diagrams representing the polarization of the
1283: atomic core by external fields.
1284: The dashed loop is the Coulomb field.
1285: (The diagrams we have presented are lowest-order exchange diagrams;
1286: there are also direct diagrams.)
1287: In diagrams (a) and (b) the core is
1288: polarized by a single field.
1289: Diagram (c) corresponds to the polarization of the core by both fields.}
1290: \label{fig:corepol}
1291: \end{figure}
1292: 
1293: \begin{figure}[b]
1294: \centerline{\psfig{file=2ndorder.eps,clip=}}
1295: \caption{Second-order correlation diagrams.
1296: Dashed line is the Coulomb interaction.
1297: Loop is the polarization of the atomic core.}
1298: \label{fig:2ndorder}
1299: \end{figure}
1300: 
1301: \begin{figure}[b]
1302: \centerline{\psfig{file=sechain2.eps, clip=}}
1303: \caption{Chaining of the self-energy operator.}
1304: \label{fig:sechain2}
1305: \end{figure}
1306: 
1307: \begin{figure}[b]
1308: \centerline{\psfig{file=screening.eps,clip=}}
1309: \caption{Screening of the Coulomb interaction.}
1310: \label{fig:screening}
1311: \end{figure}
1312: 
1313: \begin{figure}[b]
1314: \centerline{\psfig{file=hpchain.eps, clip=}}
1315: \caption{Hole-particle interaction in the polarization operator.}
1316: \label{fig:hpchain}
1317: \end{figure}
1318: 
1319: \begin{figure}[b]
1320: \centerline{\psfig{file=hpscreense.eps, clip=}}
1321: \caption{The electron self-energy operator with screening and hole-particle
1322: interaction included.}
1323: \label{fig:hpscreense}
1324: \end{figure}
1325: 
1326: \begin{figure}[b]
1327: \centerline{\psfig{file=pncdom.eps, clip=}}
1328: \caption{Lowest-order correlation corrections to the PNC E1 transition
1329: amplitude.}
1330: \label{fig:pncdom}
1331: \end{figure}
1332: 
1333: \begin{figure}[b]
1334: \centerline{\psfig{file=pncint.eps, clip=}}
1335: \caption{Small corrections to the PNC E1 transition amplitude:
1336: external field inside the correlation potential.
1337: In diagrams (a) the weak interaction is inside the
1338: correlation potential ($\delta \hat{\Sigma}$ denotes the change in
1339: $\hat{\Sigma}$ due to the weak interaction);
1340: this is known as the weak correlation potential.
1341: Diagrams (b) represent structural radiation
1342: (photon field inside the correlation potential).}
1343: \label{fig:pncint}
1344: \end{figure}
1345: 
1346: \begin{figure}[b]
1347: \centerline{\psfig{file=rad.eps, clip=}}
1348: \caption{High-frequency contribution to radiative corrections.
1349: Diagram (a) corresponds to the Uehling potential.
1350: Diagram (b) is the vertex correction.
1351: The double line represents the bound electron;
1352: the single solid line is the free electron;
1353: the Coulomb interaction is denoted by the dashed line;
1354: and the filled circle denotes the nucleus.}
1355: \label{fig:rad}
1356: \end{figure}
1357: 
1358: 
1359: \end{document}
1360: 
1361: 
1362: 
1363: 
1364: 
1365: 
1366: 
1367: 
1368: 
1369: 
1370: 
1371: 
1372: 
1373: 
1374: 
1375: 
1376: 
1377: 
1378: 
1379: 
1380: 
1381: 
1382: 
1383: 
1384: 
1385: 
1386: 
1387: 
1388: 
1389: 
1390: 
1391: 
1392: 
1393: 
1394: 
1395: 
1396: 
1397: 
1398: 
1399: 
1400: 
1401: 
1402: 
1403: 
1404: 
1405: 
1406: 
1407: 
1408: 
1409: 
1410: 
1411: 
1412: 
1413: 
1414: 
1415: 
1416: 
1417: 
1418: 
1419: 
1420: 
1421: 
1422: 
1423: 
1424: 
1425: 
1426: 
1427: 
1428: 
1429: 
1430: 
1431: 
1432: 
1433: 
1434: 
1435: 
1436: 
1437: