1: \documentclass[showpacs,amssymb,
2: %preprintnumbers,amsmath,
3: eqsecnum,
4: twocolumn,
5: tightenlines,
6: %preprint
7: ]{revtex4}
8:
9:
10:
11: \usepackage{bm}
12:
13: \usepackage{amsmath}
14:
15: \usepackage{dcolumn}
16:
17: % \usepackage[dvips]{gphicx}
18: % \sloppy
19: % \draft
20:
21: \begin{document}
22:
23: \bibliographystyle{apsrev}
24:
25: \title {Radiative corrections to parity-non-conservation in
26: atoms}
27:
28: \author{M.Yu.Kuchiev}
29: \email[Email:]{kuchiev@newt.phys.unsw.edu.au}
30: \author{V.V.Flambaum}
31: \email[Email:]{flambaum@newt.phys.unsw.edu.au}
32: \affiliation{ School
33: of Physics, University of New South Wales, Sydney 2052,
34: Australia}
35:
36: % \affiliation{
37: % School of Physics, University of New South Wales,Sydney 2052,
38: % Australia}
39:
40: % \date{\today}
41:
42: \begin{abstract} Recent progress in calculations of QED radiative
43: corrections to parity nonconservation in atoms is reviewed. The
44: QED vacuum polarization, the self-energy corrections and the
45: vertex corrections are shown to be described very reliably by
46: different methods used by different groups. All new calculations
47: have recently converged to very close final values. Each
48: separate radiative correction is very large, above 1 \% for
49: heavy atoms, but having different signs they partly compensate
50: each other. Our results for the radiative corrections for all
51: atoms are presented. The corrections are $-0.54 \% $ for
52: $^{133}$Cs, and $-0.70 \% $ for $^{205}$Tl, $^{208}$Pb, and
53: $^{209}$Bi. The result for $^{133}$Cs reconciles the most
54: accurate atomic experimental data for the $6s-7s$ PNC amplitude
55: in $^{133}$Cs of Wood {\it et al} \cite{wood_97} with the
56: standard model.
57:
58: \end{abstract}
59:
60: \pacs{32.80.Ys, 11.30.Er, 31.30.Jv}
61:
62: \maketitle
63:
64:
65: \section{Introduction}
66: \label{intro}
67:
68: It has become clear quite recently that the QED corrections give
69: a large contribution, of the order of $\sim 1 \%$, to the parity
70: nonconservation (PNC) amplitude in heavy atoms, which makes them
71: important for an analysis of modern experimental data. The present
72: paper summarizes recent progress in calculations of the QED
73: radiative corrections to PNC. Over the last few years the
74: experimental data of the Boulder group for the $6s-7s$ PNC
75: amplitude in $^{133}$Cs, which remains the most accurate in the area
76: of atomic PNC, were presumed to be able to challenge the standard
77: model. The radiative corrections eliminate this possibility, their
78: accurate determination brings the atomic experimental data for the
79: mentioned transition in agreement with the standard model. The
80: paper presents also a brief description of the methods used for
81: calculations of the QED corrections. Some of them, that have been
82: recently designed for applications to the PNC problem,
83: may be of interest in other areas where QED corrections are
84: studied.
85:
86: Several atoms and several transitions have been studied
87: experimentally in relation to PNC. As was mentioned, the most
88: accurate experimental data was obtained for the PNC $6s-7s$ PNC
89: transition in $^{133}$Cs. Studies of this transition initiated by
90: Bouchiat and Bouchiat \cite{bouchiat_74}, were continued and
91: developed in Paris
92: \cite{bouchiat_82,bouchiat_84,bouchiat_85,bouchiat_86,%
93: bouchiat_86_1,bouchiat_02} and in Boulder
94: \cite{boulder_85,boulder_86,boulder_88,wood_97}. The latest work of
95: Wood {\it et al} \cite{wood_97} reduced the experimental error to
96: $0.35\%$ that remains the best accuracy in atomic PNC experiments.
97: These results inspired interest in very accurate atomic PNC
98: calculations that, eventually, led to recognition of the fact that
99: the QED corrections contribute at the necessary level of accuracy.
100: Other experiments on PNC have been carried out for several
101: transitions in several atoms. This includes the $6p_{1/2} -
102: 7p_{1/2}$ transition in $^{205}$Tl
103: \cite{berkeley_79,berkeley_81,berkeley_85}, and the $6p_{1/2} -
104: 6p_{3/2}$ transition in $^{205}$Tl
105: \cite{oxford_91_Tl,oxford_95,seattle_95}, the $^3P_0 -^3P_1$
106: transition in $^{208}$Pb \cite{seattle_83,seattle_93,oxford_96} and
107: two transitions in $^{209}$Bi, $^4S_{3/2} - ^2D_{5/2}$ measured in
108: pioneering works of Barkov and Zolotorev
109: \cite{novosibirsk_78,novosibirsk_78_1,novosibirsk_79,novosibirsk_80}
110: and in \cite{moscow_84,oxford_87_5/2,oxford_93} as well as the
111: $^4S_{3/2} - ^2D_{3/2}$ transition
112: \cite{seattle_81,oxford_87_3/2,oxford_91_Bi}. The comprehensive
113: theoretical background on PNC as well as a detailed account of
114: events which led to a discovery of PNC in atoms (that was also the
115: first observation of PNC in neutral currents) can be found in the
116: book of Khriplovich \cite{khriplovich_91}.
117:
118: Errors of final quantities that can be extracted from experimental
119: results on PNC in atoms crucially depend on the accuracy of atomic
120: calculations. The most difficult and cumbersome part of the
121: theoretical research is the atomic structure calculations that take
122: into account the many-electron nature of the atomic wave function
123: in heavy atoms. A foundation for the modern theoretical
124: description of many-electron correlations in the PNC problem was
125: laid out in the works \cite{dzuba_89,blundell_90} which also give
126: references to previous papers. The calculations of the
127: many-electron correlations in the $6s-7s$ transition in $^{133}$Cs
128: in these works gave very close results, with the error in both sets
129: of calculations estimated as $\sim 1 \% $. The problem was
130: revisited in several recent publications that supported the results
131: of \cite{dzuba_89,blundell_90} improving the error. Improvement of
132: the theoretical error was initiated in Refs.
133: \cite{bennett_wieman_99} that compared a number of experimentally
134: measured quantities, such as polarizabilities, dipole amplitudes
135: and hyperfine constants, with theoretical predictions of Refs.
136: \cite{dzuba_89,blundell_90}, concluding that the theoretical error
137: of these works was probably better than it had been anticipated, at
138: the level of $0.4 \%$. This conclusion allowed the authors of
139: \cite{bennett_wieman_99} to reveal a deviation that existed between
140: the experimental data of \cite{wood_97} and the standard model
141: ($\sim 2 \sigma $). The following theoretical works
142: \cite{kozlov_01,dzuba_01,dzuba_02} demonstrated that indeed, the
143: theoretical error that originates from the manyelectron
144: correlations can be reduced down to $\sim 0.5 \% $ for the $6s-7s$
145: transition in $^{133}$Cs, thus matching the experimental accuracy
146: of Ref. \cite{wood_97}. The net result of the experimental and
147: theoretical progress at this period of time was a pronounced
148: deviation, of the order of $2 \sigma$, between predictions of the
149: standard model and experimental data that was proved to be not
150: related to uncertainties imposed by the complicated manyelectron
151: nature of the problem.
152:
153: The progress outlined above indicated that either the standard
154: model needed improvement, or some unrecognized, underestimated
155: factors should come into play. A careful analysis revealed that
156: there was an overlooked factor well within the framework of the
157: standard model, namely the conventional QED corrections to the PNC
158: amplitude. Obviously, they have always been kept in mind, but for
159: a long period of time they were estimated as negligible in the PNC
160: problem. From the perspective of the present day situation the
161: fact that they play an important role can be considered as a
162: general statement. Historically, however, its realization stemmed
163: from several works devoted to different, separate phenomena that
164: contribute to the QED corrections. Derevianko \cite{derevianko_00}
165: found that the Breit corrections to electron-electron interaction
166: in heavy atoms are much larger than it had been anticipated, giving
167: $-0.6 \% $ for the $6s-7s$ PNC amplitude in $^{133}$Cs. This
168: result was obtained numerically, and later confirmed in a number of
169: papers \cite{dzuba_harabati_01,kozlov_01, sushkov_01} both
170: numerically and analytically. The account of the Breit corrections
171: put the experimental data of \cite{wood_97} much closer to the
172: standard model, but this was only the beginning of an intriguing
173: zigzag road of research. An analysis of the Breit corrections by
174: Sushkov \cite{sushkov_01} revealed that one can expect the QED
175: vacuum polarization to be as important as the Breit corrections.
176: Johnson {\it et al} \cite{johnson_01} for the first time proved
177: that indeed, the vacuum polarization is very important. For the
178: $6s-7s$ PNC amplitude in $^{133}$Cs atom they found that it gives
179: $0.4 \%$, as confirmed in
180: \cite{milstein_sushkov_01,dzuba_01,kf_jpb_02}. From a purely
181: pragmatic, numerical point of view the vacuum polarization mostly
182: compensated for the contribution of the Breit corrections, bringing
183: the experimental data for the $^{133}$Cs atom back in contradiction
184: with the standard model.
185:
186: The most difficult and, correspondingly, challenging part of the
187: QED corrections to the PNC amplitude is presented by the QED
188: self-energy corrections. We are using this term here in a sense
189: that embraces the vertex corrections as well. One could have
190: anticipated these corrections to be large from the very beginning.
191: The self-energy corrections are known to give a large contribution
192: to a number of phenomena in atomic physics, including the Lamb
193: shift, the hyperfine interaction, and the energy shift due to the
194: finite nuclear size (FNS). The latter phenomenon will play a
195: particular role in the following discussion. There is no physical
196: reason indicating that the PNC amplitude should be any different.
197: The sign of the self-energy corrections is also predetermined. The
198: QED self-energy is known to be mostly repulsive (this is definitely
199: true for the $s_{1/2}$ state that should play a very important
200: role) in contrast to the attractive vacuum polarization. Therefore
201: the self-energy corrections should make the electron wave function
202: smaller at the nucleus and, consequently, give a negative
203: contribution to the PNC amplitude.
204:
205: However, the aforementioned reasons were confronted by
206: counter-arguments indicating that the self-energy corrections
207: should be small. One reason, briefly mentioned in \cite{johnson_01},
208: relies on the fact that in the plane-wave approximation
209: the self-energy corrections are small, of the order of $0.1 \% $
210: for $^{133}$Cs atom, as was found by Marciano and Sirlin
211: \cite{marciano_sirlin_83} and Lynn and Sandars
212: \cite{lynn_sandars_94}. One can argue that account of the
213: nuclear Coulomb field should not produce any dramatic enhancement,
214: and the self-energy corrections should, probably, remain small.
215: The argument has its merit, it is known to be correct for the case
216: of the vacuum polarization, where higher-order corrections are
217: suppressed, see section \ref{vacuum and related}. However, the
218: self-energy corrections behave differently. The well-studied case
219: of the hyperfine interaction is an example. It is known from
220: both analytical and numerical calculations, see Refs.
221: \cite{blundell_97,sunnergren_98} and references therein, that in
222: this problem the manifestations of the Coulomb field in the
223: self-energy corrections in heavy atoms are more pronounced than a
224: contribution of the plane-wave approximation, the latter one is
225: given by the known $Z$-independent Schwinger term $\alpha/2\pi $,
226: while the former rise quickly with the nuclear charge
227: \footnote{This example for some time was part of a theoretical
228: folklore. One of us (M.K.), who was a newcomer in the field of
229: PNC, heard it first from M.G.Kozlov. At that time the validity of
230: this argument could not probably be fully appreciated because
231: another folklore wisdom claimed that the PNC should be different
232: from everything studied previously due to a different operator
233: structure, which is not the case, as is well established now, see
234: Fig.\ref{four}.}.
235:
236: Another case, that looked even stronger, against a possible role of
237: the self-energy corrections in the PNC problem was made by Milstein
238: and Sushkov in Ref.\cite{milstein_sushkov_01}. They presented
239: results of direct analytical calculations claiming that the
240: self-energy corrections are positive and small, concluding that the
241: experimental data on the $6s-7s$ PNC amplitude in the $^{133}$Cs
242: atom are in contradiction with the standard model. The impact of
243: this work was strong, for a long period of time it was considered as
244: a reliable argument indicating that the experimental data can
245: challenge the standard model. The opposite opinion expressed in Ref.
246: \cite{dzuba_01}, which presented a model for rough estimation of the
247: self-energy corrections, was not considered as convincing, probably
248: because the proposed model neglected the vertex corrections and was
249: not gauge invariant.
250:
251: The analytical calculations of Ref.\cite{milstein_sushkov_01} seemed
252: to prove that the self-energy corrections are small and positive,
253: while some physical arguments mentioned above indicated
254: otherwise. To find the way out of this controversy we proposed in
255: Ref. \cite{kf_prl_02} an approach that avoided difficult
256: analytical calculations. This work derived a new equality, we
257: will call it the chiral invariance identity due to reasons
258: explained in section \ref{equality}. It expressed the self-energy
259: corrections for the PNC amplitude in terms of the self-energy
260: corrections to the energy shifts due to FNS. The latter ones were
261: calculated numerically in Refs.
262: \cite{johnson_soff_85,blundell_92,%
263: blundell_sapirstein_johnson_92,cheng_93,lindgren_93,cheng_02}.
264: Thus the found equality allowed one for the first time to find the
265: accurate self-energy corrections for the PNC amplitude for heavy
266: atoms. As one could have anticipated, they were large and
267: negative. In particular, for the $6s-7s$ transition in${133}$Cs we
268: found $-0.73(20)\% $ that brought the experimental data of Wood
269: {\em et al} in agreement with the standard model. The uncertainty
270: of this result was mainly due to errors of the numerical results
271: of Ref. \cite{cheng_93} that we used as input data. The most
272: pronounced among them is the error for the self-energy corrections
273: to the energy shift induced by the FNS in the $p_{1/2}$ wave. The
274: authors of \cite{cheng_93}, as well as others in the relevant
275: papers above, estimated their results as very reliable for heavy
276: atoms $Z\sim 90$, and less accurate for lighter atoms, making
277: numerical calculations below $Z=55$ difficult.
278:
279: In order to verify and refine our results, there were performed
280: analytical calculations of the leading, linear term in the $Z
281: \alpha $-expansion for the self-energy (and vertex) corrections
282: for the PNC amplitude \cite{k_jpb_02}. The result of this work
283: confirmed that of \cite{kf_prl_02}, the self-energy corrections
284: were found to be large and negative, in clear contradiction with
285: Ref. \cite{milstein_sushkov_01}. Combining the data of
286: \cite{kf_prl_02}, which is accurate for heavy atoms, with the
287: analytical data of \cite{k_jpb_02} that is reliable for light
288: atoms, Ref. \cite{k_jpb_02} found the self-energy corrections for
289: the PNC amplitude for all atoms. This analysis also allows one to
290: reduce the impact of the above mentioned errors of the data
291: extracted from \cite{cheng_93}. The result for the $6s-7s$
292: transition in $^{133}$Cs was $-0.9 \% $, in good agreement with
293: $-0.73(20) \% $ of \cite{kf_prl_02}. These two works, using
294: different techniques, unambiguously demonstrated that the
295: experimental data of Wood {\it et al} \cite{wood_97} support the
296: standard model.
297:
298: The contradiction that existed between this statement and
299: conclusions of Ref. \cite{milstein_sushkov_01} was, fortunately,
300: eradicated by the the work of Milstein {\it et al}
301: \cite{milstein_sushkov_terekhov_02} that appeared shortly after
302: \cite{k_jpb_02}. Milstein {\it et al} arrived at results that are
303: close to ours \cite{kf_prl_02,k_jpb_02}. For the $6s-7s$
304: transition in $^{133}$Cs they found the self-energy corrections to
305: be $-0.85\% $. A careful comparison of the results of the two
306: groups presented in Ref. \cite{kf_comm_02} revealed a similar
307: agreement for all values of $Z$. In this latter work it was also
308: proposed that the remaining small discrepancy is due to
309: limitations of accuracy in calculations of Milstein {\it et al}
310: that are based on $Z\alpha$-expansion ($Z\alpha = 0.4$ for Cs).
311: This conclusion has recently been proved true by Milstein {\it et
312: al} who refined their calculations in
313: \cite{milstein_sushkov_terekhov_Log_all_orders}. Now, for a vast
314: area of the nuclear charges $Z$ the results of Milstein {\it et
315: al} for the self-energy corrections agree with ours very
316: closely, the remaining deviation is below $9 \% $. The deviation
317: manifests itself more prominently for large $Z$, and we believe
318: that it is still due to those terms of the $Z \alpha $-expansion
319: that remain unaccounted for by the analytical calculations of
320: \cite{milstein_sushkov_terekhov_Log_all_orders}. Thus, using
321: several different techniques, the two teams arrived to one and the
322: same result for the self-energy corrections. The fact that the
323: results of Milstein {\it et al} converged to ours after a period
324: of initial deviation makes the final results even more convincing.
325:
326:
327: The plan of this paper is to discuss in some detail the radiative
328: corrections, summarizing all available methods and data. As an
329: introduction to this study let us first present very briefly
330: several known facts about relativistic corrections in general. The
331: main relativistic correction is, obviously, the correction that is
332: due to the relativistic nature of the Dirac equation that governs
333: the electron propagation. It is usually taken care of in the
334: initial approximation that relies on the Dirac-Hartree-Fock
335: approach, using then the machinery of the many-body theory to
336: account for the electron correlations. The next necessary step is
337: to include the relativistic correction to the photon propagator
338: that describes the electron-electron interaction. In the
339: nonrelativistic approach this propagator is approximated by its
340: Coulomb part. Relativistic effects can be taken care of
341: perturbatively, using the Breit corrections for this propagator.
342: This approximate approach is convenient because the Breit
343: correction can also be incorporated in the initial Hartree-Fock
344: mean field and then dealt with by conventional many-body methods.
345: At the same time this approximation provides sufficient accuracy.
346:
347: The next layer of corrections originates from the QED radiative
348: corrections. Numerically, as well as parametrically, they may be
349: comparable with a contribution of the Breit corrections, but it is
350: convenient to consider them separately. The radiative corrections in
351: the lowest order of the perturbation theory are
352: divided into two classes. One of them constitutes the QED vacuum
353: polarization, another one consists of the self-energy and the vertex
354: corrections. These latter two corrections are often grouped together
355: and called the e-line corrections.
356:
357: The discussion below focuses entirely on the QED radiative
358: corrections. This presumes that every other above mentioned
359: relativistic correction has been reliably taken care of in the
360: previous research on the atomic PNC problem, see the references discussed
361: above. The present paper demonstrates that {\it all} radiative
362: corrections have recently been taken into account and calculated
363: reliably. Thus, combining the radiative corrections with the previous
364: theoretical data one can be absolutely certain that {\it every}
365: possible relativistic correction in many-electron atoms is accounted
366: for. This done, the theory comes up with an important conclusion
367: \cite{kf_prl_02,k_jpb_02}: the most accurate available experimental
368: data for parity nonconservation of Wood {\it et al} \cite{wood_97}
369: fully support the standard model.
370:
371:
372:
373:
374: %\section{ Manyelectron correlations }
375:
376: %\section{ Breit corrections }
377:
378: \section{ Vacuum polarization and related phenomena }
379: \label{vacuum and related}
380:
381: \subsection{Variation of the electron wave function due to short-range
382: potentials}
383: \label{wave function}
384:
385: The effective Hamiltonian $H_{\mathrm {PNC}}$ induced by the
386: $Z$-boson exchange that describes the PNC part of the weak
387: interaction between the atomic electrons and the nucleus can be
388: presented in the following form (the relativistic units $\hbar = c
389: =1,~e^2=\alpha$ are used, if not stated otherwise)
390:
391:
392: \begin{equation}\label{05} H_{\mathrm {PNC}} = \frac{1}{ 2 \sqrt{2}
393: }\, G_{\mathrm F} \, Q_{\mathrm W}\, \rho(r)\,\gamma_5 ~.
394: \end{equation}
395: Here $G_{\mathrm F}$ and $ Q_{\mathrm W} $ are the Fermi constant
396: and the nuclear weak charge, and $\rho(r)$ is the nuclear density.
397: The PNC amplitude describes the matrix element that mixes two
398: states of opposite parity for a valence atomic electron. According
399: to Eq.(\ref{05}) the electron wave function in this matrix element
400: must be taken on the nucleus. Therefore the largest PNC amplitude
401: arises due to the mixing of the $s_{1/2}$ and $p_{1/2}$ partial
402: waves $ \langle \psi_{n,s,1/2}| H_{\mathrm {PNC}} | \psi_{n',p,1/2}
403: \rangle $, here $n,n'$ refer to the electron energy states, the
404: electron wave functions are the Dirac spinors
405:
406: \begin{equation}\label{spinor}
407: \psi_{nlj}({\bf r}) =
408: \left(
409: \begin{array}{cc} f_{nlj}(r)\Omega_{jl\mu}({\bf n}) \\
410: i g_{nlj}(r)\Omega_{j\tilde l \mu}({\bf n})
411: \end{array} \right)~,
412: \end{equation}
413: $f_{nlj}(r)$ and $g_{nlj}(r)$ are the large and small radial
414: components of the spinor, $\Omega_{j l \mu}({\bf n})$ and
415: $\Omega_{j\tilde l \mu}({\bf n})= -({\mbox {\boldmath
416: $\sigma$}}\cdot {\bf n})$ $\Omega_{j l m}({\bf n})$ are the
417: spherical spinors, and $l+\tilde l =2 j$.
418:
419:
420:
421: In applications the electron wave function is known in some
422: reasonably good approximation (say, the Hartree-Fock-Dirac initial
423: approximation plus contributions of the many-electron
424: correlations). Inevitably there exist a number of perturbations
425: that are not included in this approximation and can affect the
426: electron wave function. The QED vacuum polarization is one of such
427: perturbations, some others are considered below. In many cases the
428: perturbation is localized in the vicinity of the nucleus, at
429: distances $r$ that are much smaller than the Bohr radius of the
430: K-shell $r \ll 1/(\alpha Z m)$. This is definitely the case for the
431: QED vacuum polarization localized at $r \le 1/m$. For this range of
432: distances the behavior of the electron wave function is quite simple
433: because it is governed mostly by the pure Coulomb field of the
434: nucleus. One can anticipate therefore that the influence of the
435: perturbation on the electron wave function can be presented in
436: simple terms, directly via the perturbative potential. This
437: anticipation proves true, there exists a simple way to account for
438: the perturbation that was suggested in Ref.\cite{kf_jpb_02}.
439:
440: Note that it suffices to find the relative variation of the
441: electron wave function induced by a perturbation. Moreover, for
442: small distances $r \ll 1/(\alpha Z m)$ the relative variation does
443: not depend significantly on $r$. Therefore it is sufficient to find
444: the relative variations of the wave function at the origin, i. e.
445: to find $\delta f_{nlj}(0)/f_{nlj}(0)$ and $\delta
446: g_{nlj}(0)/g_{nlj}(0)$, where $\delta f_{nlj}(r), \delta g_{nlj}(r)
447: $ are the variations of the large and small components of the wave
448: function due to the considered perturbation, and the relative
449: variations at the origin are taken in the limit $r\rightarrow 0$. The
450: limit is necessary because either the large or small components of
451: the wave function vanish at the origin, see Eq.(\ref{bc}) below.
452: Ref. \cite{kf_jpb_02} found for the relative variations of the
453: wave functions the following result
454:
455:
456: \begin{equation}\label{final}
457: \frac{\delta f(0)}{f(0)}=\frac{\delta g(0)}{g(0)}=
458: -\frac{m}{\hbar^2}\,\int_0^\infty V(r)\,(a+kr)\,{\rm d}r~.
459: \end{equation}
460: Here and below the indices $nlj$ for the wave functions are
461: suppressed to simplify notation, $a$ is a parameter with length
462: dimension and $k$ is a dimensionless coefficient
463:
464: \begin{eqnarray} \label{a}
465: a &=& \frac{Z\alpha}{\gamma}\,\frac{\hbar}{mc}~,
466: \\ \label{k}
467: k &=& \frac{2\kappa(2\kappa-1)}{\gamma(4\gamma^2-1)}~,
468: \end{eqnarray}
469: $V(r)$ is the perturbative potential. Parameters $\gamma, \kappa$
470: are defined conventionally $\gamma = (1-(Z \alpha )^2)^{1/2}$ and
471: $\kappa = l(l+1)-j(j+1)-1/4 = \pm (j+1/2)$. Relations
472: (\ref{final}),(\ref{a}) are presented in absolute units, to make
473: them more convenient for different applications below. The simple,
474: clear formula (\ref{final}) solves the problem formulated above,
475: presenting a variation of the wave function at the origin in
476: transparent terms, over the two first moments of the potential. It
477: is instructive to consider Eq.(\ref{final}) in the nonrelativistic
478: limit that reads
479:
480: \begin{equation}\label{nonrelat}
481: \frac{ \delta \psi( 0 ) }{ \psi(0) } =
482: -\frac{m}{\hbar^2} \,\frac{2}{2l+1}\int_0^\infty V(r)\,r\,{\rm d}r~.
483: \end{equation}
484: Here $\psi(r)$ is the Schroedinger nonrelativistic wave function.
485: Deriving Eq.(\ref{nonrelat}) from (\ref{final}) one uses the fact
486: that according to (\ref{a}) the parameter $a$ goes to zero in the
487: limit $Z\alpha \rightarrow 0$, while from (\ref{k}) one finds $k
488: \rightarrow 2/(2l+1)$. There is a simple short-cut derivation
489: that leads to (\ref{nonrelat}). One expresses the variation of the
490: Schroedinger electron wave function via the Green function $G$
491:
492: \begin{equation}\label{delta}
493: \delta \psi(0) = \int G({\bf 0},{\bf r}')\,V(r')\,
494: \psi({\bf r}')\,{\rm d} ^3 r'~,
495: \end{equation}
496: approximating the Green function by the Green function for free
497: motion $G^{(0)}$
498:
499: \begin{eqnarray}\label{G0} G({\bf r},{\bf r}') &\rightarrow & G^{(0)}_l
500: ({\bf r},{\bf r}') \simeq
501: \\ \nonumber
502: &-&\frac{2m}{\hbar^2} \frac{1}{2l+1}
503: \frac{r_<^l}{r_>^{l+1}} \sum_{m=-l}^{l} Y_{lm}({\bf n})
504: Y_{lm}^\dag({\bf
505: n}') ~,
506: \end{eqnarray}
507: where $G^{(0)}_l({\bf r},{\bf r}')$ is the Green function for the
508: free motion in the $l$-th partial wave, ${\bf n} = {\bf r}/r,~ {\bf
509: n}' = {\bf r}'/r'$, and the last identity in Eq.(\ref{G0}) takes
510: into account the fact that the binding energy of the valence
511: electron is negligible for short distances. Remembering that for
512: nonrelativistic motion the wave function behaves like
513: $\psi_{l}({\bf r}) \propto Y_{lm}({\bf n})\,r^l,~r \rightarrow 0$
514: one immediately derives Eq.(\ref{nonrelat}) from Eq.(\ref{delta}).
515:
516: The point to be noted in this simple derivation is the dominance of
517: the kinetic term in the nonrelativistic limit. For small distances
518: the nonrelativistic kinetic energy behaves as $\propto 1/r^2$,
519: being larger than the potential energy. That is why Eq.(\ref{G0})
520: holds in the nonrelativistic limit. In the relativistic approach
521: the situation is different. The kinetic term behaves as $\propto
522: 1/r$, while the potential behaves as $Z \alpha /r$. One has to
523: expect henceforth that the contribution of the Coulomb potential
524: energy to the final answer is to be suppressed compared with the
525: contribution of the kinetic term only by a factor $Z \alpha $.
526: This is exactly what happens in Eq.(\ref{final}). The second term
527: $\propto kr$ on the right-hand side can be considered as a mere
528: modification of the nonrelativistic result, while the first one
529: $\propto a$ is the expected new term proportional to $Z \alpha $.
530: Thus the structure of Eq.(\ref{final}) is predetermined. The
531: calculations of Ref. \cite{kf_jpb_02} establish the coefficients
532: in this relation as well as the dependence on higher powers of the
533: $Z \alpha $-expansion that appear through the parameter $\gamma$.
534:
535: The nonrelativistic Eq.(\ref{nonrelat}) shows that the parameter
536: that governs the variation of the wave function is $\int m
537: V(r)r{\rm d}r$ that is almost an obvious result valid for a
538: variety of quantum mechanical problems \cite{LLIII}. It is
539: interesting that relativistic Eq.(\ref{final}) shows that there
540: exists another relativistic parameter $\int ma V(r){\rm d}r$. It
541: is suppressed compared with the nonrelativistic one only by a
542: factor $Z\alpha$ that is not small for heavy atoms. This
543: suppression can be well compensated for if the potential increases
544: at small distances which makes $ \int ma V(r){\rm d}r$ larger than
545: $m \int V(r)r{\rm d}r$. In this case the new relativistic
546: parameter becomes dominant, as the examples below demonstrate. The
547: singular rise of the potential at small distances is typically
548: smeared out by the finite nuclear size. Dealing with such singular
549: potentials it is important to keep in mind that the first term on
550: the right-hand side of Eq.(\ref{final}) is proportional to $Z$.
551: Therefore it originates from those distances where the full
552: strength of the electric field created by the nuclear charge is
553: present, i. e. from the region outside the nuclear core $r \ge
554: r_{\mathrm N}$. Inside the nucleus the Coulomb field created by
555: the nuclear charge diminishes, making this region less important.
556: In order to estimate its contribution, one can introduce for
557: distances $r \le r_{\mathrm N}$ an effective nuclear charge $Z
558: \rightarrow Z_\mathrm{ eff}= (r/r_\mathrm{N} )^3 \,Z$ that
559: diminishes with the nuclear radius taking into account the
560: decline of the electric field inside the nucleus.
561: Correspondingly modifying the parameter $a$ in Eq.(\ref{final})
562:
563: \begin{equation}
564: \label{Zeff}
565: a \rightarrow a_\mathrm{ eff}= \left(
566: \frac{r}{ r_\mathrm{N} }\right)^3 \,a~,
567: \end{equation}
568: one can apply this equation for the region $r\le r_{\rm N}$. We
569: will see that this approximate procedure gives sensible results
570: for the vacuum polarization, see Eqs.(\ref{w})-(\ref{w2}).
571:
572: Let us apply Eq.(\ref{final}) to the PNC problem, in which one
573: needs to calculate the PNC matrix element $ \langle
574: \psi_{n,s,1/2}| H_{\mathrm {PNC}} | \psi_{n',p,1/2} \rangle$. Both
575: $s_{1/2}$ and $p_{1/2}$ wave functions are influenced by the
576: perturbation $V(r)$. Therefore the relative variation of the
577: PNC amplitude can be found simply by adding variations of
578: $s_{1/2}$ and $p_{1/2}$ states given in Eq.(\ref{final})
579:
580: \begin{eqnarray}\label{weak}
581: \delta_{\mathrm {PNC},\,V} &\equiv & \frac{\delta
582: \langle \psi_{s,\,1/2}|H_{\mathrm {PNC}} |\psi_{p,\,1/2}\rangle}
583: {\langle \psi_{s,\,1/2}|H_{\mathrm {PNC}} |\psi_{p,\,1/2}\rangle} =
584: \\ \nonumber
585: &&-\frac{2m}{\hbar^2}\,\int_0^\infty V(r)\,\left(a+\frac{2}{3}
586: kr\right)\,{\rm d}r~.
587: \end{eqnarray}
588: Here $\delta \langle \psi_{s,\,1/2}|H_{\mathrm {PNC}}
589: |\psi_{p,\,1/2}\rangle$ is the variation of the PNC matrix element
590: due to variations of the wave functions created by the potential
591: $V(r)$. Deriving this result we take into account that essential
592: parameters for the $s_{1/2}$ and $p_{1/2}$ states are
593: $\kappa_{s}=-1,~\kappa_{p}=1, ~\gamma_s = \gamma_p =
594: (1-(Z\alpha)^2)^{1/2}\equiv \gamma$ and define $k$ in (\ref{weak})
595: to be $k\equiv k_s= 6/\Big( \gamma(4\gamma^2-1) \Big)$.
596:
597: It was assumed in Eq.(\ref{final}) that the perturbative potential
598: is a scalar. In applications it is necessary also to deal
599: with more sophisticated, tensor-type perturbations. The electric
600: and magnetic fields created by the nuclear dipole, quadrupole and
601: higher moments give examples of such perturbations. It is
602: interesting that Eq.(\ref{final}) can be generalized to describe
603: the tensor-type perturbations in simple clear terms. Suppose that
604: we have the valence electron described by the Dirac spinor
605: $\psi_{lj}({\bf r})$ (the energy of the valence electron is so low
606: that we can neglect it, suppressing the irrelevant index $n$).
607: Suppose that there is some tensor perturbation $V({\bf r})$. When
608: we consider the PNC amplitude we need to evaluate the wave
609: function at the nucleus. There are only two partial waves that
610: remain large at the origin, $s_{1/2}$ and $p_{1/2}$. This means
611: that in the multipole expansion of the $V({\bf
612: r})\psi_{lj}({\bf r})$ term in the Dirac equation only these
613: two partial waves contribute. The angular structure of these
614: partial waves is predetermined, compare Eq.(\ref{spinor}),
615: therefore we may write
616:
617: \begin{equation}
618: \label{partial}
619: [V({\bf r})\psi_{lj}({\bf r})]_{L,1/2} =
620: \left(
621: \begin{array}{cc} F_{L,1/2}(r)\Omega_{1/2,L,\mu}({\bf n}) \\
622: i \,G_{L,1/2}(r)\Omega_{1/2,\tilde L, \mu}({\bf n})
623: \end{array} \right)~,
624: \end{equation}
625: where $[V({\bf r})\psi_{lj}({\bf r})]_{L,1/2}$ means the
626: admixture of the $L_{1/2}$ state in the term $V({\bf
627: r})\psi_{lj}({\bf r})$. Eq.(\ref{partial}) shows that the
628: contribution of the term $V({\bf r})\psi_{lj}({\bf r})$ to the
629: Dirac equation is expressed in terms of the scalar functions
630: $F_{L,1/2}(r)$ and $G_{L,1/2}(r)$ with $L=0,1$. We can
631: introduce now some {\it effective} potentials $V^{\mathrm
632: {eff}}_L(r)$, $W^{\mathrm {eff}}_L(r) $ that satisfy the
633: following conditions
634:
635: \begin{eqnarray} \label{effect1} F_{L,1/2}(r) &=& V^{\mathrm
636: {eff}}_L(r) f_{L,1/2}(r)~, \\ \label{effect2} G_{L,1/2}(r) &=&
637: W^{\mathrm {eff}}_L(r) g_{L,1/2}(r)~.
638: \end{eqnarray}
639: Here $f_{L,1/2}(r),~ g_{L,1/2}(r)$ are the known large and small
640: components of the nonperturbed wave function for the valence
641: electron in the $L_{1/2}$ partial wave. The procedure described
642: shows that the term $V({\bf r})\psi_{lj}({\bf r})$ in the Dirac
643: equation at small distances can be described by the effective
644: potentials $V^{\mathrm {eff}}_L(r)$ and $ W^{\mathrm {eff}}_L(r)$
645: (which can be different for the large and the small components
646: for an arbitrary perturbation). Correspondingly, the variation
647: of the wave function $\delta \psi_{lj}({\bf r})$ at small
648: distances can be expressed via its multipole expansion in which
649: only the $s_{1/2}$ and $p_{1/2}$ states are retained
650:
651: \begin{equation}
652: \label{wf,partial}
653: \delta \psi_{lj}({\bf r}) \simeq
654: \sum_{L=0,1}
655: \left(
656: \begin{array}{cc} \delta f_{lj;\, L,1/2}(r)
657: \Omega_{1/2,L,\mu}({\bf n}) \\
658: i \delta g_{lj;\, L,1/2}(r)\Omega_{1/2,\tilde L, \mu}({\bf n})
659: \end{array} \right)~.
660: \end{equation}
661: Here the functions $\delta f_{lj;\, L,1/2},~\delta g_{lj;\,
662: L,1/2}$ describe the admixture of the $L_{1/2}$ states to the
663: function $\psi_{lj}({\bf r})$ that arise due to the tensor-type
664: perturbation. They are to be found from the Dirac equation. We
665: see that the complicated multipole problem is reduced to the
666: problem with the scalar potential, the only distinction is that
667: the role of the scalar potential plays the effective potentials
668: $V^{\mathrm {eff}}_L(r),~W^{\mathrm {eff}}_L(r)$ defined above.
669: However, this distinction makes no difference, we introduce the
670: effective scalar potentials in the Dirac equation and solve it
671: regardless of the nature of the potentials. We can rely
672: on the discussion in \cite{kf_jpb_02} to derive formulas
673: similar to Eq.(\ref{final}). The result reads
674:
675: \begin{eqnarray}\label{eff1}
676: \frac{\delta f_{lj;\, L,1/2}(0)}{f_{L,1/2}(0)}&=&
677: -\frac{m}{\hbar^2}\int_0^\infty
678: \!\!\!\!V^{\mathrm {eff}}_{\mathrm L}(r)
679: (a+kr){\rm d}r,
680: \\ \label{eff2}
681: \frac{\delta g_{lj;\, L,1/2}(0)}{g_{L,1/2}(0)} &=&
682: -\frac{m}{\hbar^2}\int_0^\infty
683: \!\!\!\!W^{\mathrm {eff}}_{\mathrm L}(r)
684: (a+kr){\rm d}r.
685: \end{eqnarray}
686: Remember that here $\delta f_{lj;\, L,1/2}(0),~\delta g_{lj;\,
687: L,1/2}(0)$ describe the admixture of $L_{1/2},~L=0,1$ states
688: that arise in the wave function $\psi_{lj}({\bf r})$ of the
689: valence electron in the state with the quantum numbers $lj$ under
690: the influence of the tensor-type perturbation $V({\bf r})$, while
691: $f_{L,1/2}(r),~g_{L,1/2}(r)$ describe the wave function of the
692: valence electron in the state with the quantum numbers $L_{1/2}$.
693: The latter ones are the same as the ones used in
694: Eqs.(\ref{effect1}),(\ref{effect2}) to define the effective
695: potentials. This makes our results presented in
696: Eqs.(\ref{eff1}),(\ref{eff2}) independent of the normalization of
697: the wave functions $f_{L,1/2}(r),~g_{L,1/2}(r)$. In
698: particular, they are independent of the energy of the valence
699: electron (as they should be since this energy is too low to produce
700: any significant effects for small distances). The quantum numbers
701: in Eqs.(\ref{a}),(\ref{k}) that define the parameters $a,k$ in the
702: right-hand sides of Eqs.(\ref{eff1}),(\ref{eff2}) correspond to
703: the state $L_{1/2},~L=0,1$. We see that for an arbitrary
704: tensor-type potential the result remains very simple. The only
705: modification, compared with the simplest scalar case, is that one
706: needs first to define the effective potentials using Eqs.
707: (\ref{effect1}),(\ref{effect2}).
708:
709:
710: Summarizing, we described in this section simple useful
711: procedures that express the variation of the valence electron wave
712: function at the origin directly in terms of the perturbative
713: potential.
714:
715:
716:
717:
718:
719: \subsection{QED vacuum polarization}
720: \label{vacuum}
721:
722: The vacuum polarization in the lowest order of perturbation
723: theory is described by the Uehling potential $V_{\mathrm VP}$
724: \cite{uehling}
725:
726: \begin{eqnarray}\label{uehling}
727: V_{\rm VP}(r) &=&
728: -\frac{2\alpha}{3\pi} \,\left(\frac{Ze^2}{r}\right)\,
729: \\ \nonumber
730: &\times& \int_1^\infty e^{-2mr\zeta}Y(\zeta)\,{\rm d}\zeta, \\ \label{Y}
731: Y(\zeta) &=& \left(1+\frac{1}{2\zeta^2}\right)
732: \frac{ \sqrt{ \zeta^2-1}}{\zeta^2}.
733: \end{eqnarray}
734: Importantly, the Uehling potential (\ref{uehling}) is singular at
735: the origin, its asymptotic for $mr \ll 1$ reads
736:
737: \begin{equation}
738: \label{uo}
739: V_{\mathrm {VP}}(r) = - \frac{2 \alpha}{3\pi} \, \left( \frac{Ze^2}{r}
740: \right) \, \left( \ln\frac{1}{mr}-C-\frac{5}{6}\right),
741: %\quad m r\ll 1~,
742: \end{equation} the singularity is smeared out by the nuclear finite
743: size $r_{\mathrm N}$. In Eq.(\ref{uo}) $C = 0.577\dots$ is the
744: Euler constant. The function $\ln mr $ arises due to conventional
745: scaling of the QED coupling constant $e^2$ that manifests itself at
746: short distances. Alternatively, this fact is referred to as the
747: unscreening of the nuclear charge at small distances. For large
748: distances $mr \gg 1~$ the Uehling potential exponentially decreases
749:
750: \begin{equation}
751: \label{exp} V_{\rm VP}(r) = - \left( \frac{Ze^2}{r} \right) \,
752: \frac{\alpha \, \exp(-2mr) }{ 4 \pi^{1/2} \, (mr)^{3/2} }~.
753: %\quad \quad mr \gg 1~.
754: \end{equation}
755: In order to find the influence of the vacuum polarization on the PNC
756: amplitude we substitute Eq.(\ref{uehling}) into Eq.(\ref{weak}).
757: Integration over the variable $r$ is performed analytically
758: resulting in the following expression for the relative correction
759: to the PNC amplitude
760:
761: \begin{eqnarray}\label{w}
762: \delta^{(\mathrm {VP}) } &=&
763: \delta^{( \mathrm{VP},\,1) } +
764: \delta^{(\mathrm{VP},\,2)},
765: \end{eqnarray}
766: where
767:
768: \begin{eqnarray}
769: \label{w1}
770: \delta^{ ( \mathrm {VP},\,1 ) } &=&
771: \frac{Z \alpha^2}{\gamma} \left(
772: \,\frac{3}{4}\,\,\frac{1}{4\gamma^2-1} + \right.
773: \\ \nonumber
774: &&
775: \left.
776: Z \alpha \,\frac{4}{3\pi}
777: \int_1^\infty E_1(2 mr_N \zeta) Y(\zeta) \,{\mathrm d}\zeta
778: \right)~,
779: \\ \label{w2}
780: \delta^{( \mathrm {VP,\,2) } } &=&
781: - \frac{ Z \alpha }{ 2\gamma } \,
782: ( \,V_{ \mathrm{VP} } ( r_{\mathrm N} ) \, r_{ \mathrm N }\,)~.
783: \end{eqnarray}
784: Here the function $E_1(x)$ is defined conventionally
785: \begin{equation}
786: \label{E1} E_1(x) = \int_1^\infty \exp(-xt)\,\frac{dt}{t}~.
787: \end{equation}
788: Eq.(\ref{w1}), which was first found in \cite{kf_jpb_02}, gives
789: the main contribution \footnote { Eqs.(43),(45) of
790: \cite{kf_jpb_02} missed a factor of 2 in the argument of the
791: $E_1$ function. This resulted in an overestimation of the role
792: of the vacuum polarization, that was claimed to be $0.47 \% $
793: for the $^{133}$Cs atom, while the present paper predicts $0.40
794: \% $.}. Deriving it one restricts the integration of
795: the singular term $ \int m a V_\mathrm{VP}$ in Eq.(\ref{weak}) to
796: the region outside the nucleus $r\ge r_\mathrm{N}$. The large
797: Coulomb field of the nucleus makes this region important. Inside
798: the nucleus $r\le r_\mathrm{N}$ the electric Coulomb field
799: diminishes, suppressing the contribution of this region. In order
800: to estimate the contribution of small distances $r\le
801: r_\mathrm{N}$ we use a modification of Eq.(\ref{weak}) introducing
802: into Eq.(\ref{final}) the effective parameter $a$ defined in
803: Eq.(\ref{Zeff}). The resulting integral in Eq.(\ref{final}) is
804: saturated in the nuclear interior region in the vicinity of the
805: nuclear surface, where the electric field remains large. This
806: fact allows one to take the values of $\gamma$ and
807: $V_\mathrm{VP}(r)$ directly at the nuclear surface. The described
808: procedure results in estimate Eq.(\ref{w2}).
809:
810: Eqs.(\ref{w})-(\ref{w2}) give the weak interaction matrix element
811: for an arbitrary atom in a transparent analytical form without
812: fitting parameters. Numerical results are easily obtained by a
813: straightforward one-dimensional integration in Eq.(\ref{w1}). One
814: only needs to specify the nuclear size that can be taken
815: conventionally as $r_N = 1.1 A^{1/3}$ fm, where $A$ is the
816: atomic number \footnote{The coefficient in the relation $r_N =
817: \eta \cdot A^{1/3}$ fm is usually taken in the region $\eta =
818: 1.1 - 1.2$. Fixing its value at $\eta =1.1$ one makes a choice,
819: though the variation in the mentioned region produces a small
820: effect.}. Alternatively, the right-hand side of Eq.(\ref{w}) can
821: be calculated using an expansion in powers of $mr_N \ll 1$ that
822: reads \cite{kf_jpb_02}
823:
824:
825: \begin{eqnarray}
826: \label{log2}
827: \!\!\!\!&&\delta^{( \mathrm {VP} ) } =
828: \frac{\alpha}{\gamma} \left\{
829: \,\frac{3}{4(4\gamma^2-1)}\,Z\alpha
830: + \right.
831: \\
832: \nonumber
833: \!\!\!\!&&
834: \left.
835: \frac{2}{3\pi}(Z\alpha)^2\left[
836: \left(\ln\frac{1}{mr_N}-C-\frac{5}{6}\right)^2 +
837: 0.759\right] \right\},
838: \end{eqnarray}
839: where the omitted terms are $O(mr_{\mathrm N})$ \footnote{This
840: expansion brings (\ref{w}) to a form that is close, but not
841: identical to the one derived in \cite{milstein_sushkov_01}. We do
842: not pursue the origin for this discrepancy since calculations in
843: the cited paper were restricted by the logarithmic accuracy.}.
844: The error of simplification of Eq.(\ref{weak}) to (\ref{log2})
845: remains $ \le 7 \% $ for heaviest atoms, decreasing for lighter
846: atoms. A very interesting feature revealed by Eq.(\ref{log2}) is
847: the second power of the logarithmic factor in the term $\sim
848: Z^2 \alpha^3 \ln^2(m r_{\mathrm N})$. This strong, double-logarithm
849: enhancement was first found by Milstein and Sushkov
850: \cite{milstein_sushkov_01}. The derivation presented above makes
851: the origin of this effect clear. The first power of the logarithmic
852: function $\ln m r$ originates from the unscreening of the nuclear
853: charge at small separations (\ref{uo}), while the second one
854: appears from the relativistic parameter $ \int V(r)
855: dr\propto \int dr \ln (mr)/r\propto \ln ^2 1/m r_\mathrm{N} $ in
856: Eq.(\ref{weak}).
857:
858:
859: In Fig. \ref{one} we present the relative corrections due to the QED
860: vacuum polarization $\delta^{(\mathrm{VP})}$ to the PNC matrix
861: element calculated with the help of Eqs.(\ref{w})-(\ref{w2}). The
862: corrections are positive and rapidly grow for heavy atoms. For the
863: most interesting case of the $^{133}$Cs atom Eq.(\ref{w}) gives
864: $\delta^{(\mathrm{VP})} =0.40 \% $. For $^{205}$Tl, $^{208}$Pb,
865: and $^{209}$Bi Eq.(\ref{w}) predicts corrections
866: $\delta_{(\mathrm{VP})} = 0.93,~0.96,~0.99\% $ respectively. For
867: all four atoms the nuclear interior region gives 10 \% of the total
868: correction, $\delta^{(\mathrm{VP},\,2)} = 0.1
869: \delta^{(\mathrm{VP})} $, which means that to make results accurate
870: numerically this region needs to be taken care of.
871:
872: %Note also that $\delta_{\mathrm{VP},\,2}$ substantially varies with
873: % $Z$.
874: %
875:
876: Compare these results with other results reported recently.
877: Johnson, Bednyakov and Soff \cite{johnson_01} calculated for the
878: first time the correction due to the vacuum polarization for the
879: $6s-7s$ PNC amplitude in $^{133}$Cs and found it to be $0.4\% $. The
880: result of \cite{johnson_01} includes, along with the variation of
881: the weak matrix element, variations of the dipole matrix element
882: and the corresponding energy denominators that, combined together,
883: describe the $s_{1/2}-s_{1/2}$ amplitude measured experimentally.
884: Dzuba {\it et al} \cite{dzuba_01} confirmed this result,
885: calculating the correction $0.41 \% $, and supplied more details
886: providing separate variations for all three quantities mentioned
887: above. They found that variations of the dipole matrix elements and
888: the energy denominators induced by the Lamb shift corrections,
889: being not small, compensate each other almost completely. Thus the
890: variation of the PNC amplitude proves to be equal to the variation of
891: the PNC matrix element. The numerical result for thallium found in
892: \cite{df_03} gives a correction $0.94 \% $, to be compared with
893: $0.93\% $ mentioned above. Calculations of Milstein {\em et
894: al} in Ref.\cite{milstein_sushkov_01} were restricted by the
895: logarithmic accuracy that was further improved by using some
896: constant as a fitting parameter to reproduce $0.4\% $ for
897: $^{133}$Cs, in line with \cite{johnson_01}. For Tl
898: Ref.\cite{milstein_sushkov_01} finds $0.9 \% $. This is slightly
899: less than the above mentioned prediction of Eq.(\ref{w}) ($0.93 \%
900: $). This (insignificant) discrepancy probably arises
901: because the procedure adopted in \cite{milstein_sushkov_01} does
902: not account for the increase of the contribution of the nuclear
903: interior region with the nuclear charge. We conclude that all
904: reported calculations for the vacuum polarization correction are in
905: very good agreement, distinctions between results derived by
906: different methods and by different groups are all below $5 \%$.
907:
908:
909: Let us discuss now possible corrections to the results discussed
910: above. One of them originates from a modification of the vacuum
911: polarization due to the finite nuclear size. The corresponding
912: generalization of the Uehling potential was given by Fullerton
913: and Rinker \cite{fullerton_rinker}. In order to estimate its
914: contribution we need to remember that the main $\propto \ln^2
915: mr_\mathrm{N}$ term in Eq.(\ref{log2}) originates from the
916: integration well outside of the nuclear radius where the Uehling
917: potential is very close to the one of Fullerton and Rinker.
918: Therefore this main part of the correction is not changed.
919: Modifications could arise only for the integration in the vicinity
920: of the nuclear surface, but this region by itself gives only a
921: fraction ($\sim 10 \% $) of the total correction, while the
922: relative difference of the Uehling potential and the Fullerton and
923: Rinker potential on the nuclear surface is small. For all atoms it
924: remains $\le 5 \% $, as we found by direct numerical
925: comparison of the Uehling potential with the Fullerton and Rinker
926: potential. We conclude that the modification of the vacuum
927: polarization due to the finite nuclear size is irrelevant.
928:
929:
930: Another possible modification can originate from the QED
931: higher-order corrections that contribute to the vacuum
932: polarization. One of them stems from the fact that the Landau
933: pole requires that the linear in $\ln mr$ term in
934: Eq.(\ref{uo}) be followed by higher-order logarithmic
935: terms. However, the corresponding series runs in powers of $Z
936: \alpha(\alpha \ln mr)^n, ~n=1,2 \dots$, giving
937: small contributions for higher order terms $n>1$.
938:
939: Consider now a modification of the Uehling potential that is due to
940: the fact that the Coulomb field of the nucleus disturbs the virtual
941: electron-positron pair. This effect is conveniently described with
942: the help of the Wichmann-Kroll potential \cite{wichmann-kroll}
943: that, compared to Eq.(\ref{uehling}), describes the next term in
944: the $Z \alpha $-expansion. For estimates it is sufficient
945: to take the Wichmann-Kroll potential at small distances, where its
946: asymptotic $mr\ll 1$ reads \cite{milstein_strakhovenko_83}:
947:
948: \begin{equation}\label{wk}
949: V_{\rm WK}(r) = 0.092 \,\frac{2 \alpha}{3\pi}\,(Z \alpha)^2 \, \left
950: ( \frac{Ze^2}{r} \right)~. \end{equation} Using
951: Eq.(\ref{weak}) one immediately finds from Eq.(\ref{wk}) the ratio of
952: the correction induced by the Wichmann-Kroll potential to the Uehling
953: correction
954:
955: \begin{equation}\label{ratio}
956: \frac{ \delta ^ \mathrm{ ( VP ) } _ \mathrm {WK} }
957: { \delta^\mathrm{ (VP) }_\mathrm{ U } } =
958: -0.184 \,
959: \frac{ ( Z \alpha )^2 } { \ln \Big( 1/(mr_\mathrm{N} ) \Big) }~,
960: \end{equation}
961: as was found in \cite{kf_jpb_02}. For $^{133}$Cs this ratio is
962: about $-0.007 $, demonstrating that the correction due to the
963: Wichmann-Kroll potential is very small. Thus all higher order QED
964: terms in the vacuum polarization are negligible.
965:
966: Alongside the vacuum polarization there exists also another
967: polarization effect, namely the polarization of the nucleus. One
968: can estimate its influence on the electron wave functions and
969: the PNC matrix element with the help of the conventional
970: polarization potential $V_\mathrm{NP}$ that is created by the
971: nucleus
972:
973: \begin{equation}
974: \label{nucPol}
975: V_\mathrm{NP} (r) = - \frac{ e^2\, \alpha_\mathrm{d} }
976: { 2 r^4 }~.
977: \end{equation}
978: Here $\alpha_\mathrm{ d}$ is the static nuclear dipole
979: polarizability. It can be estimated roughly assuming that the total
980: oscillator strength for the nuclear dipole excitations is located
981: at the energy of nuclear giant dipole resonance
982: $\Omega_\mathrm{d}$. Then, using the conventional sum rule for the
983: oscillator strengths one finds
984:
985: \begin{equation}
986: \label{DipRes}
987: \alpha_\mathrm{d} \simeq \frac{ Z e^2 }
988: { m_p \, \Omega_\mathrm{d}^2 }~,
989: \end{equation}
990: where $m_p$ is the proton mass. Since the potential (\ref{nucPol})
991: increases at small separations the main contribution in
992: Eq.(\ref{weak}) is given by the first term in Eq.(\ref{weak}) which
993: should be integrated outside the nuclear radius $r\ge r_\mathrm { N
994: }$, as was explained above. Having this in mind, one substitutes
995: (\ref{nucPol}) in (\ref{weak}) finding an estimate for the
996: influence of the nuclear polarization on the PNC matrix element
997:
998: \begin{equation}
999: \label{PNCnucPol}
1000: \delta^\mathrm{ (NP) } \simeq
1001: \frac{2}{3}\, \frac{Z^2 \alpha^3}
1002: {\gamma \,m_p\,\Omega_ \mathrm { d }^2 \,r_\mathrm{N}^3 } \,~.
1003: \end{equation}
1004: For the $^{133}$Cs atom we take $\Omega_\mathrm {d} \simeq 15$ MeV
1005: obtaining $\delta^{(\mathrm{ NP })} \simeq 0.01 \% $ that shows that
1006: the nuclear polarization is negligible. This conclusion agrees with
1007: estimates of Milstein and Sushkov \cite{milstein_sushkov_01}, who
1008: proposed that the effect is of the order of $\sim 0.1
1009: Z^{2/3}\alpha^2$ \footnote{ Eq.(12) of \cite{milstein_sushkov_01}
1010: suggests that the corrections are negative, which is probably a
1011: typo since the static polarization inevitably results in an
1012: increase of the effect.}, which numerically is close
1013: to Eq.(\ref{PNCnucPol}).
1014:
1015:
1016: Summarizing, we demonstrated above that the QED vacuum
1017: polarization gives a large positive correction ($\sim 1 \% $) to
1018: the PNC amplitude. Contributions of higher order QED corrections
1019: are much smaller. This means that the influence of the nuclear
1020: Coulomb potential on the virtual electron-positron pair is not
1021: pronounced, possibly due to opposite signs of the virtual
1022: particles. We will see below that for a single electron the
1023: situation is different, the Coulomb field does play an important
1024: role for the self-energy correction, see Eq.(\ref{197}). The
1025: influence of the finite nuclear size on the vacuum polarization
1026: was found to be nonessential, the polarization can be well
1027: described by the simplest Uehling potential. Also negligible is
1028: the polarization of the nucleus. We presented above in some
1029: detail analytical methods of calculation, paying less attention
1030: to purely numerical methods that have played a very important role
1031: in the problem, see \cite{johnson_01,dzuba_01,df_03}. Omissions
1032: of details in our presentation of numerical methods is justified
1033: by the fact that the vacuum polarization is described by a
1034: sufficiently simple potential, which can be incorporated in the
1035: initial Hartree-Fock-Dirac approximation for the atomic
1036: electrons, and then dealt with by conventional methods of
1037: many-body theory. Thus the numerical approach follows
1038: the mainstream of atomic structure calculations discussed in some
1039: detail elsewhere, see for example \cite{dzuba_89,blundell_90}.
1040: Overall, the polarization is a well understood, robust
1041: effect that is well described by Eqs.(\ref{w})-(\ref{w2}).
1042:
1043:
1044: \section{ Self-energy and vertex corrections }
1045:
1046:
1047: \subsection{Chiral Invariance and QED corrections }
1048: \label{equality}
1049: Let us show, following Ref.\cite{kf_prl_02}, that there exists an
1050: interesting and useful relation that gives the QED radiative
1051: corrections to the PNC matrix element via similar radiative
1052: corrections to another quantity, the energy shifts of the atomic
1053: electron induced by the finite nuclear size (FNS). This relation can
1054: be written as
1055:
1056: \begin{equation}\label{ddd}
1057: \delta_{\mathrm {PNC},\,sp} = \frac{1}{2}\,
1058: ( \delta_{ \mathrm {FNS},\,s } +
1059: \delta_{\mathrm {FNS},\,p })~,
1060: \end{equation}
1061: where $\delta_{ \mathrm {PNC},\, sp }$ is the relative radiative
1062: correction to the PNC matrix element between the $s_{1/2}$ and
1063: $p_{1/2}$ orbitals
1064:
1065: \begin{equation}\label{d} \delta_{\mathrm {PNC},\,sp} =
1066: \frac{\langle \psi_{s,\,1/2} | H_{\mathrm {PNC}}
1067: | \psi_{p,\,1/2}\rangle^ {{\mathrm {rad}} } } { \langle
1068: \psi_{s,\,1/2} | H_{\mathrm {PNC}} |\psi_{p,\,1/2}\rangle~~~ }~.
1069: \end{equation}
1070: The energy difference between the considered opposite parity
1071: orbitals $\sim 1$ eV is much lower than a typical excitation energy
1072: $\sim m=0.5$ MeV that governs the radiative corrections. We can
1073: therefore neglect this difference assuming that $E_{s,1/2}\simeq
1074: E_{p,1/2}$. This assumption makes the correction $\langle
1075: \psi_{s,\,1/2} | H_{\mathrm PNC}|\psi_{p,\,1/2}\rangle^{\mathrm rad}$
1076: gauge invariant. To make this argument more transparent, one can
1077: use the Coulomb approximation for the atomic field and consider the
1078: degenerate Coulomb $ns_{1/2}$ and $np_{1/2}$ levels.
1079:
1080: The problem of gauge invariance can be viewed from another
1081: perspective. One can consider the full PNC
1082: amplitude. Alongside the PNC matrix element it includes
1083: also the conventional E1 amplitude and the corresponding
1084: energy denominator. The PNC amplitude describes the events on
1085: the mass shell, the photon frequency in the E1 transition
1086: equals the excitation energy of the atom. Therefore the
1087: amplitude is definitely gauge invariant. Considering the
1088: radiative corrections in the lowest order, one can separate
1089: them into corrections to the PNC matrix element, the Lamb-shift
1090: corrections to the energies, and the corrections to the E1
1091: transition. Their sum is definitely gauge invariant. The point
1092: is that the latter two corrections are known to compensate
1093: each other almost completely \cite{dzuba_01}. The only
1094: remaining corrections are the ones which are associated with
1095: the PNC matrix element. Thus these latter corrections are
1096: gauge invariant. To make this discussion complete, let us
1097: mention also that in the perturbation theory approach outlined
1098: in the next section \ref{Zalpha} the gauge invariance of the
1099: PNC matrix element is self-evident. There is no doubt,
1100: therefore, that one can assume that the radiative corrections to the
1101: PNC matrix element are gauge invariant.
1102:
1103: The quantities $\delta_{ {\mathrm {FNS}},\,s}$ and
1104: $\delta_{{\mathrm {FNS}},\,p}$ in Eq.(\ref{d}) are the
1105: relative radiative corrections to the FNS energy shifts
1106: $E_{\mathrm {FNS},\,s}, ~E_{\mathrm {FNS},\,p},$ for the
1107: chosen $s_{1/2}$ and $p_{1/2}$ electron states
1108:
1109: \begin{eqnarray}\label{dd} \delta_{ {\mathrm {FNS}},\,l} =
1110: \frac{ E_{ {\mathrm {FNS}},\,l}^{\mathrm {rad} } }
1111: { E_{\mathrm {FNS},l } } ~, \quad l = 0,1 ~. \end{eqnarray}
1112: The FNS energy shifts can be presented as matrix elements of
1113: the potential $ V_{\mathrm {FNS}}(r)$ that describes the deviation
1114: of the nuclear potential from the pure Coulomb one inside the
1115: nucleus due to the spread of the nuclear charge
1116:
1117: \begin{equation}
1118: \label{EFNS}
1119: E_{{\mathrm {FNS}},\,l } =
1120: \langle \psi_{ l,\,1/2 } | V_{\mathrm {FNS}}
1121: |\psi_{l,\,1/2}\rangle~,\quad l=0,1.
1122: \end{equation}
1123: Equality (\ref{ddd}) may be established for the sum of all QED
1124: radiative corrections ($\sim Z \alpha^2 f(Z \alpha)$), or
1125: specified for any gauge invariant class of them. We will discuss
1126: first the self-energy and vertex corrections, restricting our
1127: consideration to the lowest order of perturbation theory
1128: presented by the Feynman diagrams in Fig.\ref{two}, calling them
1129: the e-line corrections. Later we will switch our attention to
1130: the vacuum polarization and demonstrate that it satisfies
1131: (\ref{ddd}) as well.
1132:
1133: The intermediate electron states in diagrams of Fig. \ref{two}
1134: are described by the propagator $ G = (\gamma_\mu p^\mu +
1135: \gamma_0 U - m)^{-1}$, with $p^\mu = (\epsilon, -i {\mbox
1136: {\boldmath $\nabla$} } )$, where $\epsilon=m+E\simeq m$ is the
1137: relativistic electron energy, $U=U(r)$ is the atomic potential
1138: that includes the potential created by the nucleus with FNS and,
1139: generally speaking, the screening potential of atomic electrons.
1140: The latter one is not important at small distances, but it still
1141: can be accounted for by the formalism. For the relativistic
1142: propagator we use the same symbol $G$ that was applied
1143: previously for the nonrelativistic Green function in
1144: Eq.(\ref{delta}), which should not produce confusion. At small
1145: distances the exchange potential is negligible, therefore the
1146: potential is local. The external legs of the diagrams describe
1147: the wave functions $ \psi_{s,1/2}({\bf r})$ and
1148: $\psi_{p,1/2}({\bf r})$ for the considered $s_{1/2}$ and
1149: $p_{1/2}$ levels. Let us note first that in the region of short
1150: distances the following relations hold
1151:
1152: \begin{eqnarray}\label{prop}
1153: \gamma_5 G & = & - G \gamma_5 ~,
1154: \\ \label{wf}
1155: \psi_{s,1/2}({\bf r}) & = & ~~\,c \gamma_5 \psi_{p,1/2}({\bf r})~.
1156: \end{eqnarray}
1157: They follow from the chiral invariance that manifests itself at
1158: small distances. For $r \ll
1159: 1/m$ the mass term in the Dirac equation becomes smaller than the
1160: kinetic term ($\propto 1/r$). For even smaller distances $ r\le
1161: Z\alpha/m$ the potential ($Z\alpha/r$) also exceeds the mass
1162: term. Therefore for small distances we can neglect the mass
1163: $m\Rightarrow 0$, using the resulting chiral invariance for our
1164: advantages.
1165: Since Eqs.(\ref{prop}),(\ref{wf}) will be important for us, we
1166: examine them in some detail. In order to prove Eq.(\ref{prop})
1167: one neglects the mass term in the electron propagator assuming
1168: that
1169:
1170: \begin{equation} \label{m=0} G^{-1} = \gamma_\mu p^\mu + \gamma_0
1171: U~. \end{equation} Then, the identity $\gamma_5 \gamma_\mu = -
1172: \gamma_\mu \gamma_5$ leads to $G^{-1} \gamma_\mu = - \gamma_\mu
1173: G^{-1}$ that results in (\ref{prop}). In order to verify
1174: Eq. (\ref{wf}) let us write down the conventional Dirac equations
1175: for the large $f$ and small $g$ components of the electron wave
1176: function (\ref{spinor})
1177:
1178: \begin{eqnarray}
1179: \label{dirac1}
1180: f' +\frac{1+\kappa}{r} f - (\epsilon + m - U) g &=& 0~,
1181: \\ \label{dirac2}
1182: g' +\frac{1-\kappa}{r} g + (\epsilon - m - U) f &=& 0~.
1183: \end{eqnarray}
1184: Here we omitted the indexes $njl$, distinguishing the states by the
1185: parameter $\kappa$ that takes the values $-1$ and $1$ for the
1186: $s_{1/2}$ and $p_{1/2}$ states respectively. For small separations
1187: one can neglect in
1188: Eqs.(\ref{dirac1}),(\ref{dirac2}) the mass term $m$. After that we
1189: observe that the equations become invariant under a substitution
1190: \begin{eqnarray}
1191: \nonumber
1192: \kappa &\rightarrow& -\kappa~,
1193: \\ \ \label{1-1}
1194: f &\rightarrow& ~~\,g~,
1195: \\ \nonumber
1196: g &\rightarrow& -f~.
1197: \end{eqnarray}
1198: Remember now that according to Eq.(\ref{spinor}) the
1199: $s_{1/2}$ and $p_{1/2}$ wave functions have the form
1200:
1201: \begin{eqnarray} \label{sp}
1202: \psi_{s,1/2} =
1203: \left(
1204: \begin{array}{cc} f_{s,1/2}\,u \\
1205: i \, g_{s,1/2}(- \mbox{\boldmath{$\sigma$}} \cdot {\bf n} ) \,u
1206: \end{array} \right), \\ \nonumber
1207: \psi_{p,1/2} =
1208: \left(
1209: \begin{array}{cc} f_{p,1/2}
1210: (- \mbox{\boldmath{$\sigma$}} \cdot {\bf n})\,u \\
1211: i \,g_{p,1/2}\,u
1212: \end{array} \right)~.
1213: \end{eqnarray}
1214: Here the two component spinor $u = u_\alpha,~\alpha=1,2$ depends
1215: on the projection of the total momentum $\mu$, $u_\alpha =
1216: \delta_{\alpha,1}$ for $\mu=1/2$, and $u_\alpha =
1217: \delta_{\alpha,2}$ for $\mu=-1/2$. It follows from Eq.(\ref{sp})
1218: that
1219:
1220: \begin{equation}
1221: \label{g5p}
1222: \gamma_5 \,\psi_{p,1/2} = i
1223: \left( \begin{array}{cc} g_{p,1/2} \,u \\
1224: -i f_{p,1/2} (-\mbox{\boldmath{$\sigma$}} \cdot {\bf n}) \,u
1225: \end{array} \right)~,
1226: \end{equation}
1227: where we took into account that in the standard
1228: representation considered for the Dirac matrixes
1229:
1230: \begin{equation}
1231: \label{g5def}
1232: \gamma_5 = \left( \begin{array}{cc} 0 & 1\\ 1 &0 \end{array}
1233: \right)~.
1234: \end{equation}
1235: From Eq.(\ref{g5p}),(\ref{sp}) one observes that
1236: Eq.(\ref{wf}) can be presented as
1237:
1238: \begin{eqnarray} \label{fg} f_{s,1/2}&=& ~~\,c g_{p,1/2}~, \\
1239: \label{gf} g_{s,1/2} &=& -c f_{p,1/2}~.
1240: \end{eqnarray}
1241: We see that these latter relations coincide with
1242: Eq.(\ref{1-1}). The first substitution $\kappa \rightarrow
1243: -\kappa$ in (\ref{1-1}) exchanges equations for the $s_{1/2}$ and
1244: $p_{1/2}$ states, while the two others are identical to
1245: Eqs.(\ref{fg}) and (\ref{gf}) respectively. Thus the verified
1246: above Eqs.(\ref{1-1}) prove validity of Eq.(\ref{wf}).
1247:
1248: At this point one may recall a subtlety, a symmetry of the Dirac
1249: equation revealed by Eqs.(\ref{1-1}) is, generally speaking, not
1250: sufficient. One needs to verify also that the
1251: boundary conditions satisfy same symmetry as well. However, one
1252: can argue that the boundary conditions arise as a physical
1253: complement to the Dirac equation and, therefore, the symmetry
1254: of the equation should automatically result in the symmetry of the
1255: boundary conditions. To verify this argument one can recall that
1256: for the finite nuclear charge distribution the boundary conditions
1257: at the origin read simply
1258:
1259: \begin{equation}
1260: \label{bc}
1261: \left\{ \begin{array}{ll}
1262: f_{s,1/2}(0) = const~, & \\
1263: g_{s,1/2}(0) = 0~,
1264: \end{array} \right.
1265: \quad
1266: \left\{ \begin{array}{ll}
1267: f_{p,1/2}(0) = 0~, & \\
1268: g_{p,1/2}(0) = const~,
1269: \end{array} \right.
1270: \end{equation}
1271: where $ const \ne 0$. Eqs.(\ref{bc}) explicitly demonstrate that
1272: indeed, the boundary conditions are invariant under the
1273: substitution Eq.(\ref{1-1}). This completes our verification of
1274: Eq.(\ref{wf}). The constant $c$ in (\ref{wf}), which depends on
1275: the normalization conditions, will be irrelevant for us because we
1276: are interested in relative quantities that appear in
1277: Eq.(\ref{ddd}) where this constant will be canceled out.
1278:
1279: It is instructive to look at numerical verifications of
1280: Eq.(\ref{wf}) shown in Fig.\ref{three}. Supporting the proof
1281: given above, they provide also details that are cumbersome for
1282: analytic treatment. Fig.\ref{three} (a) shows that
1283: Eq.(\ref{fg}), which according to the discussion above holds at
1284: small distances $r\ll 1 /m$, remains approximately valid at
1285: sufficiently large distances, up to (and even beyond) the
1286: Compton radius $r \le 1/m$. In contrast, Eq.(\ref{gf}) is
1287: applicable only at small $r \ll 1/m$ as shown in Fig.
1288: \ref{three} (b), being violated at $ r \sim 1/m $, as one can
1289: see from Fig. \ref{three}(a). We can remember, however, that the
1290: dominant contribution to the PNC matrix element arises from the
1291: $f_{s,1/2}$ and $ g_{p,1/2}$ functions, which are both large for
1292: $r\le 1/m$, much larger than the other two functions
1293: $g_{s,1/2},f_{p,1/2}$. Since these large functions satisfy
1294: Eq.(\ref{fg}) even at sufficiently large distances $r \le 1/m$,
1295: one can also rely in this region on Eq.(\ref{wf}). We will
1296: exploit this opportunity at some later stage below, but our
1297: prime interest is related to the distances as small as the
1298: nuclear radius $r \le r_\mathrm{N}\ll Z\alpha/m$, where we
1299: verified beyond doubt that Eqs.(\ref{prop}),(\ref{wf}) hold, see
1300: Fig.\ref{three} (b).
1301:
1302: Consider the diagrams (b),(c) in Fig. \ref{two} that describe the
1303: self-energy correction for the PNC amplitude. The thick lines in
1304: this figure represent the electron propagation in the atomic
1305: field, the dot vertex is the PNC interaction described by
1306: Eq.(\ref{05}), and the dashed line shows the photon. Let us
1307: specify that the left and right legs of these diagrams describe
1308: the $s_{1/2}$ and $p_{1/2}$ states respectively. The PNC
1309: interaction (\ref{05}) is located at small distances $r_\mathrm{
1310: N} \ll 1/m $. Let us use this fact and, relying on
1311: Eqs.(\ref{prop})(\ref{wf}), transform the PNC matrix element. To
1312: simplify the notation it is convenient to present the Hamiltonian of
1313: the weak interaction Eq.(\ref{05}) in the form
1314:
1315: \begin{equation}\label{005} H_{\mathrm {PNC}} = V_\mathrm{PNC}
1316: (r)\,\gamma_5 ~,
1317: \end{equation}
1318: where $V_\mathrm{PNC} (r) = G_{\mathrm F} \, Q_{\mathrm W}\,
1319: \rho(r)/(2 \sqrt{2})$ is a scalar factor in the PNC interaction.
1320: Consider first the diagram (b). The $\psi_{p,1/2}({\bf r}) $
1321: function in this diagram enters the PNC interaction that makes its
1322: argument ${\bf r}$ localized inside the nucleus, $r\le
1323: r_\mathrm{N}$. This allows us to use Eq.(\ref{wf}) deriving
1324:
1325: \begin{eqnarray}\label{se(b)}
1326: \langle \psi_{p,\,1/2} | \, \gamma_5 \,V_\mathrm{PNC}\, G \,\Sigma\,
1327: |\psi_{s,\,1/2} \rangle
1328: \\
1329: \nonumber
1330: \equiv
1331: \langle \psi_{s,\, 1/2} | \, V_\mathrm{PNC} \,G \, \,\Sigma |
1332: \psi_{s,\,1/2}
1333: \rangle ~.
1334: \end{eqnarray}
1335: Here and below we use the symbol of identity ($\equiv$) that simply
1336: means dropping the irrelevant constant $c$ introduced in
1337: Eq.(\ref{prop}) (that is canceled out when the relative contribution
1338: of the correction is considered anyway). Eq.(\ref{se(b)}) uses
1339: conventional notation, $\Sigma$ and $G$ are the self-energy
1340: operator and the electron propagator mentioned above. The
1341: right-hand side of this relation shows that we can look at the diagram
1342: (b) as the one that describes the diagonal $s_{1/2}-s_{1/2}$
1343: transition induced by an effective scalar potential
1344: $V_\mathrm{PNC}$. Consider now the same diagram (b), but for
1345: another physical quantity, for the FNS energy shift for the
1346: $s_{1/2}$ level. Its contribution to the energy reads
1347:
1348: \begin{equation}\label{se(b)fns}
1349: \langle \psi_{s,\,1/2} | \, V_\mathrm{FNS}\, G \, \Sigma \, |
1350: \psi_{s,\,1/2}
1351: \rangle ~.
1352: \end{equation}
1353: Eqs.(\ref{se(b)}),(\ref{se(b)fns}) give the contribution of one and
1354: the same diagram (b) to different quantities, namely the PNC
1355: amplitude and FNS energy shifts. Comparing them one
1356: observes their similarity. In both cases the diagram describes the
1357: $s_{1/2}-s_{1/2}$ transitions that are induced by the short-range
1358: scalar potentials. The only distinction comes from the fact that
1359: the potentials $V_\mathrm{PNC}(r)$ and $V_\mathrm {FNS}(r)$ possess
1360: different profiles.
1361:
1362: This latter difference, however, proves to be irrelevant. In order to
1363: see this we need to discuss the integral with a short-range
1364: potential that arises in the diagram. Its integrand includes the
1365: potential (i. e. $V_\mathrm{PNC}(r)$ for the PNC amplitude and
1366: $V_\mathrm{FNS}(r)$ for the FNS energy shift), the electron wave
1367: function $\psi_{s,1/2}({\bf r})$ and the electron propagator $G({\bf r},
1368: {\bf r}')$. The wave function and the propagator inside the nucleus
1369: are almost constants, see discussion of this fact below. Taking
1370: them out of the integration we observe that the considered
1371: sophisticated integral is simplified, being reduced to the
1372: integral over the potential itself $\int V_\mathrm{PNC}(r) d{\bf r}$.
1373: It is easy to see that the same integral arises in the non-perturbed
1374: matrix element. Therefore in the relative contribution of the
1375: correction this integral is canceled out, and the result does not
1376: depend on the potential at all. This conclusion is valid for any
1377: short-range potential. Thus the relative contribution of the
1378: diagram (b) proves to be one and the same for the PNC and the FNS
1379: problems. This means that the relative contribution of the
1380: diagram (b) to the PNC matrix element equals half of its relative
1381: contribution to the energy shift of the $s_{1/2}$ state due to FNS.
1382: The mentioned factor $1/2$ takes into account the fact that the
1383: energy shift is induced by the two diagrams (b) and (c) in Fig.
1384: \ref{two} that are identical for the FNS problem.
1385:
1386: Before proceeding further, let us discuss the behavior of the
1387: electron wave functions inside the nucleus which was exploited in
1388: the above derivation. Note that inside the nucleus the large
1389: functions $f_{s,1/2}(r)$ and $g_{p,1/2}(r)$ exhibit smooth, almost
1390: constant behavior, as demonstrates Fig. \ref{three}. The reason
1391: for this effect stems from the fact that inside the nucleus the
1392: electric field rapidly diminishes with the radius. The functions
1393: $g_{s,1/2}(r)$ and $f_{p,1/2}(r)$, show some variation, but it
1394: remains very mild. Assessing consequences of this smooth variation
1395: one can recall that, firstly, the integration volume $r^2dr$
1396: enhances the contribution of the nuclear surface region where
1397: these functions are very smooth indeed, as shows Fig. \ref{three}
1398: (b). Secondly, their role is not prominent anyway, simply because
1399: they are smaller than the functions $f_{s,1/2}(r)\,g_{p,1/2}(r)$.
1400: Therefore they can be well approximated by constants. Thus both
1401: $s_{1/2}$ and $p_{1/2}$ states are well described by constant-type
1402: functions inside the nucleus. This fact was used above for the
1403: $s_{1/2}$ state, the argument is repeated below for the $p_{1/2}$
1404: state. We also presumed above that the propagator $G({\bf r},
1405: {\bf r}')$ does not depend on the variable ${\bf r}$, when $r\le
1406: r_\mathrm{N}$. This statement follows from the simple fact that
1407: the coordinate ${\bf r}'$ gives the location of the point where
1408: the radiation of the photon takes place, therefore $r'\sim 1/m \gg
1409: r$. The smallness of $r$ compared with $r'$ justifies its neglect
1410: in the propagator.
1411:
1412:
1413:
1414: Examining the diagram Fig.\ref{two} (c) we use the same arguments
1415: as we exploited above for the (b) diagram proving that its
1416: relative contribution to the PNC amplitude equals half of its
1417: relative contribution to the energy shift due to FNS of the
1418: $p_{1/2}$ state. Combining our results for the diagrams (b) and
1419: (c) we conclude that the self-energy corrections (superscript SE)
1420: described by these diagrams comply with Eq.(\ref{ddd})
1421:
1422: \begin{equation}
1423: \label{ds}
1424: \delta_{\mathrm {PNC},\,sp} ^\mathrm {(SE)} = \frac{1}{2}\,
1425: \Big( \, \delta_{ \mathrm {FNS},\,s }^\mathrm {(SE)} +
1426: \delta_{\mathrm {FNS},\,p }^\mathrm {(SE)}\, \Big)~.
1427: \end{equation}
1428: The given derivation does not use any specific gauge condition,
1429: therefore it is gauge invariant.
1430:
1431: Consider now the vertex correction (a) in Fig. \ref{two}. An
1432: analytical expression for it can be written in the conventional
1433: form
1434: \begin{eqnarray}
1435: \label{vertex}
1436: % \!\!\!
1437: % \!\!\!
1438: - i\alpha \int \!\! \frac{d^4q}{(2\pi)^4}
1439: d{\bf r} d{\bf r}'
1440: d{\bf r}_0
1441: D^{\mu\nu}(q)
1442: \\
1443: \nonumber
1444: \times
1445: e^ {i \,{\bf q}\cdot ({\bf r} -{\bf r}')}
1446: \Phi_{\mu\nu}({\bf r}, {\bf r}', {\bf r}_0)~,
1447: \end{eqnarray}
1448: where
1449: \begin{eqnarray}
1450: \nonumber
1451: \Phi_{\mu\nu}({\bf r}, {\bf r}', {\bf r}_0) &=&
1452: \bar \psi_{p,1/2}({\bf r}) \gamma_\mu
1453: G({\bf r},{\bf r}_0; \epsilon + \omega)
1454: \\ \nonumber
1455: &\times&
1456: V_\mathrm{PNC}(r_0)
1457: \gamma_0 \,\gamma_5 \,
1458: G({\bf r}_0 , {\bf r}'; \epsilon + \omega)
1459: \\
1460: \label{Phi}
1461: &\times&
1462: \gamma_\nu
1463: \psi_{s,1/2}({\bf r}')~.
1464: \end{eqnarray}
1465: Here $D^{\mu\nu}(q)$ is the photon propagator, $q^\mu = (\omega,
1466: {\bf q})$ is the four-vector of the photon momentum. The range of
1467: distances $r_{\mathrm rad}$ where the radiation processes take
1468: place is of the order of the Compton radius, $r_{\mathrm rad}\sim
1469: m^{-1}$. This ensures that $ r,r'\gg r_0$ because the PNC
1470: interaction is localized at $ r_0 \le r_\mathrm {N}$. We can
1471: simplify therefore Eq.(\ref{vertex}) factorizing the interaction
1472: over ${\bf r}_0$
1473:
1474: \begin{eqnarray}
1475: \nonumber
1476: && \int d{\bf r}_0 \,\Phi_{\mu\nu}({\bf r}, {\bf r}', {\bf r}_0)=
1477: \int V_\mathrm{PNC} (r_0)\, d{\bf r}_0 \\ \label{r0}
1478: && \times \, \bar \psi_{p,1/2}({\bf r}) \, \gamma_\mu \,
1479: G({\bf r},\,{\bf 0}; \,\epsilon + \omega) \\ \nonumber
1480: && \times \,\gamma_0 \,\gamma_5 \,
1481: G({\bf 0} , \,{\bf r}';\, \epsilon + \omega) \,\gamma_\nu \,
1482: \psi_{s,1/2}({\bf r}')~,
1483: \end{eqnarray}
1484: where we put $r_0=0$ in the arguments of the propagators. We
1485: observe that the potential arises only in a factor $\int
1486: V_\mathrm{PNC} (r_0)\, d{\bf r}_0$ that is separated from all other
1487: elements of the diagram. The same factor appears in the main term of
1488: the PNC amplitude. Therefore the relative contribution of the
1489: vertex correction does not depend on the potential at all. Remember
1490: that the same property is exhibited by the self-energy
1491: correction discussed above.
1492:
1493: The variables $r,r'$ describe the radiation processes and therefore
1494: are expected to be of the order of the Compton radius $r,r'\sim
1495: 1/m$. Let us consider the implications that arise if one
1496: presumes that the region of small separations
1497:
1498: \begin{equation}
1499: \label{rr}
1500: r,r' < \frac{1}{m}
1501: \end{equation}
1502: gives the dominant contribution to the integral in
1503: Eq.(\ref{vertex}). We can engage in this region
1504: Eqs.(\ref{prop}),(\ref{wf}). The first of them allows one to
1505: exchange the $\gamma_5$ matrix in Eq.(\ref{r0}) with the
1506: propagator $G$. One can also transpose it with the vertexes
1507: $\gamma_5 \gamma_\mu = - \gamma_\mu \gamma_5 $. Using several
1508: such transpositions we can bring the $\gamma_5 $ matrix from
1509: its central position in expression (\ref{r0}) to the side,
1510: where it hits the external wave function. After that we use
1511: Eq.(\ref{wf}) for this function. Repeating the procedure twice,
1512: one time shifting $\gamma_5$ to the left, and another one to
1513: the right we derive
1514:
1515: \begin{eqnarray}
1516: \nonumber
1517: && \!\! \bar \psi_{p}({\bf r}) \, \gamma_\mu \,
1518: G({\bf r},\,{\bf 0}) \, \gamma_0 \,\gamma_5 \,
1519: G({\bf 0} , \,{\bf r}') \,\gamma_\nu \,
1520: \psi_{s}({\bf r}') \\ \label{l}
1521: &&\!\! \equiv \bar \psi_{s}({\bf r}) \, \gamma_\mu \,
1522: G({\bf r},\,{\bf 0} ) \, \gamma_0 \,
1523: G({\bf 0} , \,{\bf r}') \,\gamma_\nu \,
1524: \psi_{s}({\bf r}') \\ \label{r}
1525: &&\!\! \equiv \bar \psi_{p}({\bf r}) \, \gamma_\mu \,
1526: G({\bf r},\,{\bf 0} ) \, \gamma_0 \,
1527: G({\bf 0} , \,{\bf r}') \,\gamma_\nu \,
1528: \psi_{p}({\bf r}') ~.
1529: \end{eqnarray}
1530: To simplify notation we dropped here the total momentum $1/2$
1531: from the indexes in the wave functions (i. e. $\psi_p \equiv
1532: \psi_{p,1/2}$ etc), and omitted the energy variable in the
1533: propagators (this variable equals $\epsilon+\omega$ for all
1534: propagators). The point of these transformation is that the
1535: $\gamma_5$ matrix disappears from the right-hand sides of
1536: Eqs.(\ref{l}),(\ref{r}). Substituting these two expressions
1537: back in Eqs.(\ref{r0}), (\ref{vertex}) we find that the relative
1538: contribution of the vertex diagram (a) to the PNC matrix element
1539: can be expressed either as a matrix element of the
1540: $s_{1/2}-s_{1/2}$ transition, or the matrix element of the
1541: $p_{1/2}-p_{1/2} $ transition. We recognize in these two matrix
1542: elements the relative contributions to the energy shifts due to
1543: FNS for the $s_{1/2}$ and $p_{1/2}$ states respectively. This
1544: latter conclusion takes advantage of the fact discussed above
1545: that the profile of the short-range potential does not influence
1546: the relative correction, one can conveniently choose the
1547: potential to be $V_\mathrm{PNC}$ or $V_\mathrm {FNS}$. This
1548: discussion demonstrates that the vertex (superscript V)
1549: corrections to the PNC amplitude and to the energy shifts due to
1550: FNS prove to be approximately equal
1551:
1552: \begin{equation}
1553: \label{V}
1554: \delta_{\mathrm {PNC},\,sp} ^\mathrm {(V)} \cong
1555: \delta_{ \mathrm {FNS},\,s }^\mathrm {(V)} \cong
1556: \delta_{\mathrm {FNS},\,p }^\mathrm {(V)}~.
1557: \end{equation}
1558: Combining Eqs.(\ref{ds}),(\ref{V}) we find that the total e-line
1559: corrections, i. e. the self-energy plus the vertex correction,
1560: satisfy Eq.(\ref{ddd}).
1561:
1562: The derivation presented above exploits Eq.(\ref{rr}) presuming
1563: that the small distances are dominant in the vertex correction.
1564: There are reasons indicating that this assumption does not put
1565: restrictions on the derived result (\ref{ddd}). As a first
1566: attempt one can try to exploit the behavior of the wave functions
1567: $f_{s,1/2},g_{p,1/2}$. Fig. \ref{three} shows that they
1568: approximately satisfy Eq.(\ref{wf}) even at sufficiently large
1569: separations, of the order of the Compton radius $ r \le 1/m$.
1570: This shows that Eq.(\ref{l}) remains approximately applicable
1571: even if the integration region includes sufficiently large
1572: separations $ r \sim 1/m$. This, in turn, demonstrates that the
1573: first equality in Eq.(\ref{V}), i. e. $ \delta_{\mathrm
1574: {PNC},\,sp} ^\mathrm {(V)} = \delta_{ \mathrm {FNS},\,s
1575: }^\mathrm {(V)}$ also remains approximately valid. This is a
1576: positive result. However, this fact alone is not sufficient.
1577: The second equality in Eq.(\ref{V}) proves to be much more
1578: sensitive to the essential integration region because the wave
1579: functions $f_{p,1/2},~g_{s,1/2}$ fail to satisfy Eq.(\ref{wf}) at
1580: $r \sim 1/m$. We need therefore to be more careful.
1581:
1582:
1583: Let us recall at this stage that the range of distances $r_{\mathrm
1584: rad}$ where the radiation processes take place depends on the
1585: chosen gauge for the electromagnetic field. It is well known that
1586: the usual length-gauge favors larger distances, while the velocity
1587: and, especially, acceleration forms, make the smaller distances
1588: contribute more to the radiation process. These known examples
1589: show that the region that contributes to the radiation
1590: process is {\em not} gauge invariant. This is important. If the
1591: radiation region is not gauge invariant, then we are able to choose
1592: a gauge in which this region is located close to the nucleus,
1593: inside the zone defined by Eq.(\ref{rr}). In this gauge the above
1594: derivation is valid and Eq.(\ref{V}) for the vertex is correct.
1595: Note that this argument does not rely on an explicit form for
1596: the necessary gauge. It suffices to acknowledge only that the
1597: radiation region is not gauge invariant. Summarizing, there exists
1598: a gauge (more accurately, a family of gauges) in which
1599: Eq.(\ref{V}) holds.
1600:
1601: Recall now that Eq.(\ref{ds}) for the self-energy remains valid
1602: in any gauge, in particular in the one discussed above. In this
1603: gauge we can combine together the self-energy and vertex
1604: corrections (\ref{ds}),(\ref{V}) proving the validity of their sum
1605: Eq.(\ref{ddd}). The three quantities in this latter relation are
1606: all well-defined, gauge invariant physical observables. The fact
1607: that Eq.(\ref{ddd}) is derived in one particular gauge means that
1608: it is valid in any gauge, as was proposed in \cite{kf_prl_02}.
1609:
1610: Let us summarize the above discussion. We argued that the
1611: calculations can be organized in such a way that all important
1612: events happen at small distances. In this region the chiral
1613: invariance holds, resulting in Eqs.(\ref{prop}),(\ref{wf}) and,
1614: consequently, in Eq.(\ref{ddd}). Thus this latter equality
1615: expresses the fundamental and simple fact. At small, nuclear
1616: distances the chiral invariance governs the problem.
1617: Consequently, Eq.(\ref{ddd}) that follows from it, can be called
1618: the chiral invariance identity.
1619:
1620: Let us estimate the accuracy of Eq.(\ref{ddd}). The above
1621: derivation used the gauge in which the radiation processes for the
1622: diagram (a) take place mostly at small distances $r_\mathrm {rad} <
1623: m^{-1}$. They should also take place outside the nucleus $r_\mathrm
1624: {N}<r_{\mathrm rad}$, as is necessary to justify our presumption
1625: that the shape of the potential inside the nucleus is irrelevant.
1626: From the last two inequalities we find that the derivation relies on
1627: a parameter $\xi = mr_{\mathrm N} \sim 0.01$. This determines the
1628: magnitude of the error (few per cent) of Eq.(\ref{ddd}).
1629:
1630:
1631: Similarly one considers the contribution of the QED vacuum
1632: polarization. Eqs.(\ref{final}),(\ref{a}) and (\ref{k}) present
1633: explicit variations for $s_{1/2}$ and $p_{1/2}$ wave functions at the
1634: origin induced by the vacuum polarization. Using these wave functions
1635: to calculate corrections to the PNC matrix element and FNS energy
1636: shifts, one immediately finds that the chiral invariance identity
1637: (\ref{ddd}) holds for the vacuum polarization as well. This fact is
1638: in line with Eq.(\ref{ds}) for the self-energy corrections. The proof
1639: of (\ref{ddd}) for the vacuum polarization stops here because one need
1640: not care about the complicated vertex corrections.
1641:
1642: Note that we do not consider above the radiative corrections of the
1643: order $\sim \alpha/\pi$ which appear in the plane wave approximation.
1644: These contributions have been included into the radiative corrections
1645: to the weak charge $Q_W$ (and the renormalization of the charge and
1646: electron mass in the case of FNS energy shifts). Correspondingly, we
1647: subtract the contribution of the plane waves from Eq.(\ref{ddd}),
1648: considering only the part of the corrections that depends on the
1649: atomic potential $\sim Z\alpha^2 f(Z\alpha)$. For heavy atoms this
1650: subtlety is insignificant numerically because the considered
1651: $Z$-dependent part of the correction is bigger than the omitted
1652: $Z$-independent one, as we will see below.
1653:
1654: The chiral invariance identity (\ref{ddd}) dives the e-line
1655: corrections to the PNC matrix element, which are difficult to
1656: calculate, in terms of the corrections to the FNS energy shifts that
1657: have been well-studied both numerically, by Johnson and Soff
1658: \cite{johnson_soff_85}, Blundell \cite{blundell_92}, Cheng {\em et
1659: al} \cite{cheng_93} and Lindgren {\em et al} \cite{lindgren_93}, and
1660: analytically, by Pachucki \cite{pachucki_93} and Eides and Grotch
1661: and Eides {\it et al} \cite{eides_97}. Ref. \cite{cheng_93}
1662: presents the e-line radiative corrections to the FNS energy shifts
1663: for $1s_{1/2}$, $2s_{1/2}$ and $2p_{1/2}$ levels in hydrogenlike
1664: ions with atomic charges $Z=60,70,80,90$. Eq. (\ref{ddd}) contains
1665: relative corrections, therefore we needed to calculate the FNS
1666: energy shifts $E_{\mathrm {FNS}}$. This was done in \cite{kf_prl_02}
1667: by solving the Dirac equation with the conventional Fermi-type
1668: nuclear distribution $\rho(r) = \rho_0 /\{ 1 + \exp [(r-a)/c] \} $.
1669: Parameters $a,c$ were taken the same as in \cite{cheng_93}, namely
1670: $a = 0.523$ fm and $c$ chosen to satisfy $R_{\mathrm {rms}} = 0.836
1671: A^{1/3} + 0.570$ fm.
1672:
1673: Using the results of \cite{cheng_93} and this calculation we
1674: obtained in \cite{kf_prl_02} the relative radiative corrections
1675: shown in Fig. \ref{four}. In order to include the interesting case
1676: $Z=55$ and to account for all values of $55 \le Z \le 90$ we used
1677: interpolating formulae presented in \cite{cheng_93}, as well as data
1678: of \cite{cheng_02}. The relative corrections for the $1s$ and $2s$
1679: levels are approximately the same size. This indicates that the
1680: radiative processes responsible for the correction take place at
1681: separations much smaller than the K-shell radius, $r \ll (Z \alpha
1682: m)^{-1}$, which is consistent with the assumption $r \le 1/m$ above.
1683: For these separations we can assume that, firstly, the screening of
1684: the nuclear Coulomb field in manyelectron atoms does not produce any
1685: significant effect, and, secondly, the relative corrections do not
1686: depend on the atomic energy level because for small separations all
1687: atomic $ns_{1/2}$-wave functions exhibit similar behavior. These
1688: arguments remain valid for the $p_{1/2}$ states as well, permitting
1689: us to presume that the results shown in Fig. \ref{four} for the
1690: $2s_{1/2}$-levels and $2p_{1/2}$-levels of hydrogenlike ions remain
1691: valid for $s_{1/2}$ and $p_{1/2}$ states of the valence electron in
1692: a manyelectron atom. We obtain the e-line radiative corrections for
1693: the PNC matrix element using the identity (\ref{ddd}) that expresses
1694: them via the found corrections to the FNS energy shifts. The found
1695: PNC corrections, presented in Fig. \ref{four} by the dotted line,
1696: are negative and large. For the $^{133}$Cs atom the correction is
1697: $-0.73(20) \%$, for Tl it is $-1.6 \% $. The result for $^{133}$Cs
1698: is important because it indicated for the first time that the
1699: radiative corrections reconcile the experimental data of Wood {\em
1700: et al} \cite{wood_97} with the standard model \cite{kf_prl_02}.
1701:
1702: A notable feature of the self-energy corrections presented in Fig.
1703: \ref{four} is their close similarity. The corrections to the FNS
1704: energy shifts for both $s_{1/2}$ and $p_{1/2}$ states, as well as
1705: for the PNC amplitude, have large negative contributions from the
1706: QED self-energy, which monotonically and very smoothly increase with
1707: the nuclear charge. This observation brings us back to the point
1708: mentioned in section \ref{intro}. The large and negative
1709: contribution of the self-energy for heavy atoms is exactly what one
1710: {\em should} expect from it. Anything else would be a surprise. In
1711: order to verify this point one can try to apply the approach outlined
1712: to some other problem, for example to the hyperfine interaction HFI
1713: which has been carefully examined previously, see detailed
1714: discussions in \cite{blundell_97,sunnergren_98} and references
1715: therein \footnote{ One of us (M.K.) is thankful to A.I.Milstein for
1716: the second reference.}.
1717:
1718: With this purpose one can try to derive the chiral invariance
1719: identity for the HFI. The problem is that the HFI has a long-range
1720: tail $\sim 1/r^3$ that is prominent in the region where the chiral
1721: invariance is violated. However, if one believes that convergence of
1722: the HFI matrix elements is fast enough, then the following equality
1723:
1724: \begin{equation} \label{hfi} \delta_{ {\mathrm {FNS}},\,s} \approx
1725: \delta_{ {\mathrm {HFI}},\,s}'~,
1726: \end{equation}
1727: that was proposed in \cite{kf_prl_02} should hold at least
1728: approximately. Here $\delta_{ {\mathrm {HFI}},\,s}'$ is the
1729: radiative correction to the HFI for $s_{1/2}$-levels, the primed
1730: notation indicates that the $Z$-independent Schwinger term
1731: $\alpha/(2\pi)$ should be excluded. For heavy atoms this subtlety
1732: is not important, since the contribution from the Coulomb-induced
1733: corrections is much stronger than the $Z$-independent Schwinger
1734: term. Fig.\ref{four} shows the e-line contribution to $\delta_{
1735: {\mathrm {HFI}},\,s}'$ that was extracted in \cite{kf_prl_02}
1736: from data of \cite{cheng_93} using interpolation for all values of
1737: $Z$ considered there. It agrees semi-quantitatively with
1738: Eq.(\ref{hfi}), the deviation is less than 33 \%. We observe again the
1739: same trend, the self-energy correction is large and negative, in
1740: accordance with the clear physical reasons mentioned in section
1741: \ref{intro}. Another positive conclusion is that chiral
1742: invariance identities similar to (\ref{hfi}) can be
1743: applicable even for those potentials that spread out of the
1744: nuclear core.
1745:
1746:
1747: The approach outlined raises two points. Firstly, it is
1748: interesting to have an estimate for the self-energy correction in
1749: simple terms, that would not appeal to sophisticated numerical
1750: calculations. Secondly, it is important to derive the self-energy
1751: corrections for lighter elements. Direct numerical calculations
1752: are difficult for them because the error rapidly increases for
1753: lighter atoms. The main source for the error is the FNS energy
1754: shift for the $p_{1/2}$ state. The lightest element for which the
1755: self-energy corrections have been calculated is Cs, where the
1756: error for the FNS energy shift in the $p_{1/2}$ wave was $\simeq
1757: 100 \% $ \cite{cheng_02}. Exactly this error results in the
1758: uncertainty of the above mentioned result $-0.73(20) \% $. For
1759: heavier atoms the numerical errors of \cite{cheng_93} rapidly
1760: decrease. Therefore the main error in our result for very heavy
1761: ($Z \ge 80$) atoms is associated with the validity of
1762: Eq.(\ref{ddd}), that was estimated above as a few percent.
1763:
1764: The estimate for the magnitude of the effect and calculations for
1765: lighter elements can be conveniently performed with the help of the
1766: $Z \alpha$-expansion considered in the next section.
1767:
1768:
1769:
1770:
1771: \subsection{ Perturbation theory in powers of
1772: $Z \alpha $ }
1773: \label{Zalpha}
1774:
1775: Let us consider the perturbation theory for the e-line
1776: corrections. In the initial approximation one describes the
1777: electron using plane waves and consequently including in the
1778: perturbation theory three types of processes. Firstly, the PNC
1779: interaction, secondly, the e-line corrections, and, thirdly, the
1780: Coulomb interaction. Taking the nuclear Coulomb field as a
1781: perturbation one can formulate the perturbation theory in powers
1782: of the nuclear charge $Z$. The parameter that governs the
1783: corresponding expansion is $\alpha (Z \alpha)^n$, where $n$ is the
1784: order of the perturbation theory. This statement follows from a
1785: conventional, well known result \cite{LLIII}. In the Coulomb
1786: problem the perturbation runs in powers of $Z e^2/\hbar v$, where
1787: $v$ is a typical velocity of the electron. For short distances $
1788: v\simeq 1$, thus $Z e^2/\hbar v = Z \alpha$. The type of expansion
1789: considered typically includes some additional large logarithmic
1790: factors, as specified below. In the lowest order of perturbation
1791: theory the simplest, linear in $Z$ correction is $\propto
1792: Z\alpha^2 $. Thus formulated problem requires calculation of the
1793: relevant coefficients that can be achieved using the corresponding
1794: Feynman diagrams.
1795:
1796: The Feynman diagrams for the e-line corrections in the first order
1797: of perturbation theory over the Coulomb field are presented in
1798: Fig. \ref{five}. The direct calculation of these diagrams for the PNC
1799: amplitude gives the following result for the relative contribution
1800: of the e-line corrections to the PNC amplitude
1801:
1802: \begin{equation} \label{197} \delta_{\mathrm{PNC},sp}^
1803: { (\mathrm{e-line},\,1)} = - 1.97 \,Z\alpha^2~.
1804: \end{equation}
1805: It was first obtained in \cite{k_jpb_02} that had an amusing
1806: history \cite{history}. The index 1 in the superscript of
1807: (\ref{197}) reminds one that this is the linear in $Z$ correction.
1808: For the $^{133}$Cs atom Eq.(\ref{197}) gives $-0.6 \% $, which is
1809: in line with the result $- 0.73(20) \%$ that follows from the chiral
1810: invariance identity (\ref{ddd}) as discussed in the previous
1811: section. An earlier attempt of Milstein and Sushkov
1812: \cite{milstein_sushkov_01} gave the coefficient $\approx \! +0.1$,
1813: instead of $-1.97$ in Eq.(\ref{197}). This distinction had
1814: important implications. Based on their result the authors of
1815: \cite{milstein_sushkov_01} claimed that there existed the
1816: contradiction between the experimental data of \cite{wood_97} and
1817: the standard model, while Eq.(\ref{197}) indicated that the
1818: experimental data agree with the model. Fortunately, the
1819: controversy was soon resolved by Milstein {\it et al} in the
1820: following Ref.\cite{milstein_sushkov_terekhov_02} that agreed with
1821: (\ref{197}) and found also an analytical expression for the
1822: coefficient in this equation that reads $-(23/4-4\log 2)\ =
1823: -1.970$. This result has also been confirmed in recent
1824: Ref.\cite{sapirstein_03}.
1825:
1826:
1827: At this point it is worth returning to the problem of gauge
1828: invariance. It is easy to demonstrate by conventional methods
1829: that the sum of all the diagrams in Fig. \ref{five} is gauge
1830: invariant. This fact should be compared with the discussion of
1831: the gauge invariance in Section \ref{equality}, see after
1832: Eq.(\ref{ddd}). However, for practical applications one often
1833: fixes the gauge because this allows one to simplify lengthy
1834: analytical calculations. This is the way the calculations were
1835: performed in
1836: \cite{k_jpb_02,milstein_sushkov_terekhov_02,sapirstein_03}.
1837: Since these works employed different gauges, the first one used
1838: the Feynman gauge while the second and third ones relied on the
1839: Fried-Yennie gauge (also called the Yennie gauge), their mutual
1840: agreement provides an additional helpful check of the validity
1841: of the final result Eq.(\ref{197}).
1842:
1843: Eq.(\ref{197}) provides a simple transparent estimate for the
1844: contribution of the e-line corrections. One point to be noted is
1845: a strong dependence of the result on the nuclear charge that
1846: makes the correction large for heavy atoms; remember the
1847: discussion of this point in Section \ref{intro}. This behavior
1848: of the self-energy correction differs qualitatively from what we
1849: saw for the vacuum polarization in section \ref{vacuum and
1850: related}, where the higher-orders in the $Z \alpha$-expansion
1851: were found small. A notable feature in Eq.(\ref{197}) is the
1852: large coefficient $\sim \! -2.0$ on the right-hand side. {\em
1853: Naively} one could expect this coefficient to be smaller, of
1854: the order of $ \sim 1/\pi$. It is interesting that a similar
1855: ``numerical enhancement'' occurs for the e-line radiative
1856: correction for the energy shift that is due to the finite
1857: nuclear size (FNS). This correction was examined analytically
1858: in Refs.\cite{pachucki_93,eides_97}, and more recently in Ref.
1859: \cite{milstein_sushkov_terekhov_02}. The result can be written
1860: as
1861:
1862: \begin{eqnarray} \nonumber
1863: \delta^{ ( \mathrm {e-line},1 )}_{\mathrm {FNS}\,2 } & = &
1864: -\left( \frac{23}{4}-4\ln 2\right) Z \alpha^2
1865: \\ \label{FNS}
1866: & = &- 2.978 \,Z\alpha^2 ~,
1867: \end{eqnarray}
1868: where the analytical expression was derived in
1869: \cite{milstein_sushkov_terekhov_02}. We see that indeed, the
1870: e-line corrections to the FNS energy shift are governed by the
1871: large coefficient $\sim \!-3.0$ in (\ref{FNS}), similar to
1872: (\ref{197}) for the PNC amplitude. Fig. \ref{six} examines this
1873: similarity in more detail. It shows data available for relative
1874: e-line corrections for the two problems mentioned above, namely
1875: for the PNC amplitude and FNS energy shifts. The linear in $Z$
1876: approximations (\ref{197}),(\ref{FNS}) (that are valid for
1877: sufficiently small values of $Z$) are compared in this figure
1878: with results that follow from numerical calculations of
1879: Ref.\cite{cheng_93}. We observe a very close quantitative
1880: similarity between corrections to PNC and FNS. In both cases the
1881: linear approximations (\ref{197}) and (\ref{FNS}) predict large
1882: negative corrections, which qualitatively agrees with results
1883: based on numerical calculations for heavy atoms. The numerical
1884: validity of the linear approximations seems to be limited by the
1885: region below $Z = 55$, for higher $Z$ they underestimate the
1886: effect. The Cs atom lies on the border, where results of the
1887: small-$Z$ and large-$Z$ approaches agree reasonably well.
1888:
1889:
1890: Numerical data used in Eq.(\ref{ddd}) incorporates an error that
1891: increases for smaller values of $Z$, see Tables III and IV of
1892: \cite{cheng_93}. In order to reduce the impact of this error we can
1893: combine together Eqs.(\ref{ddd}),(\ref{197}). One can
1894: approximate all nonlinear terms omitted in (\ref{197}) by the
1895: simplest quadratic function and choose the corresponding coefficient
1896: to reproduce the results based on numerical data for very large
1897: $Z$, $Z \sim 90$, where the numerical errors are small. This
1898: approach gives the following interpolating formula \cite{k_jpb_02}
1899: for the corrections to the PNC amplitude
1900:
1901: \begin{equation}\label{interp} \delta^{( \mathrm { e-line,\,int})
1902: }_{\mathrm {PNC}} = - 1.97 \, Z\alpha^2\, (\, 1+1.55 \,Z \alpha
1903: \,)~.
1904: \end{equation}
1905: Fig. \ref{six} shows that the data available for large-Z and small-Z
1906: regions is very smooth. Therefore {\em any} reasonable interpolation
1907: would produce a pattern close to Eq.(\ref{interp}). We can be
1908: certain therefore that (\ref{interp}) gives reliable numerical data.
1909: For the $^{133}$Cs atom Eq.(\ref{interp}) predicts $-0.9(1) \% $;
1910: compare this result with $-0.6 \%$ of (\ref{197}) and $ -0.73(20) \% $
1911: that follows from Eq.(\ref{ddd}) \footnote{ Ref. \cite{kf_prl_02},
1912: where this result was first derived, estimated the uncertainty by
1913: adopting the error $0.2 \% $ of Ref. \cite{kf_prl_02}. A more
1914: accurate analysis demonstrates that this estimate is too
1915: conservative; a more realistic one is $ 0.1 \% $ taken above.}.
1916: For the Tl atom Eq.(\ref{interp}) predicts $-1.60 \% $ that agrees
1917: with the prediction of (\ref{ddd}) discussed above (as it should,
1918: since the data of \cite{cheng_93} incorporates only very small
1919: errors for heavy atoms).
1920:
1921: It should be noted that Eq.(\ref{ddd}) does not rely on the
1922: perturbation theory in powers of $Z \alpha$, effectively
1923: including all nonlinear corrections. This makes it accurate
1924: even at large values of the nuclear charge. It should therefore
1925: be considered as a formula that provides a convenient short-cut
1926: presentation of the nonlinear result. A similar formula was
1927: suggested in \cite{k_jpb_02} to reconcile the data for the e-line
1928: corrections to the energy shifts due to FNS for the $s_{1/2}$
1929: state. The linear in $Z$ term of Eq.(\ref{FNS}) and numerical
1930: data of \cite{cheng_93} available for heavy atoms can be
1931: interpolated by
1932:
1933: \begin{equation}
1934: \label{intFNS}
1935: \delta^{( \mathrm{ e-line,\,int} )}_{\mathrm {FNS} }
1936: = - 2.978 \, Z\alpha^2\,
1937: (\, 1+0.85 \,Z \alpha \,)
1938: \end{equation}
1939: Fig. \ref{six} shows that this interpolation is reliable. This
1940: implies that for relatively light atoms $Z\simeq 55 - 75$
1941: numerical results for the $s_{1/2}$ energy corrections should be
1942: amended by slightly larger values.
1943:
1944:
1945: An alternative method to include the nonlinear corrections was
1946: developed in
1947: Refs.\cite{milstein_sushkov_terekhov_02,%
1948: milstein_sushkov_terekhov_Log_all_orders} that relies on direct
1949: analytical calculations. This approach faces a difficulty because
1950: even in the next-to-leading order $\sim Z^2\alpha^3$ the
1951: calculations become complicated. Ref.\cite{milstein_sushkov_01}
1952: showed, however, that in this order there appears a large
1953: logarithmic factor that makes the correction proportional to $\sim
1954: Z^2\alpha^3 \ln 1/mr_\mathrm{N}$. Calculations with the
1955: logarithmic accuracy, when some constant is neglected, are much
1956: more feasible. The elegant result for the thus calculated
1957: second-order correction (index 2 in the superscript) was derived
1958: in \cite{milstein_sushkov_terekhov_02}
1959:
1960: \begin{equation} \label{mstlog}
1961: \delta_{\mathrm{PNC},sp}^\mathrm{e-line,2} = -\frac{Z^2
1962: \alpha^3}{\pi} \left( \frac{15}{4}-\frac{\pi^2}{6} \right) \, \ln
1963: \frac{b}{mr_\mathrm{N} }~, \end{equation} where $b = \exp
1964: [1/(2\gamma) - C-5/6]$. The final result of
1965: \cite{milstein_sushkov_terekhov_02} is given by the sum of the
1966: linear term Eq.(\ref{197}) and the second order correction
1967: Eq.(\ref{mstlog}). It is compared in Fig. \ref{seven} with
1968: predictions of Eqs. (\ref{ddd}) and (\ref{interp}). For the sake
1969: of completeness the figure also shows the previous results of
1970: Ref. \cite{milstein_sushkov_01}. For all values of the nuclear
1971: charges the results of \cite{milstein_sushkov_terekhov_02} are
1972: close to our results derived from Eqs.(\ref{ddd}) and
1973: (\ref{interp}), as was emphasized in \cite{kf_comm_02}. Recently
1974: Milstein {\em et al}
1975: \cite{milstein_sushkov_terekhov_Log_all_orders} refined their
1976: arguments, demonstrating that the logarithmic factor similar to
1977: the one in Eq.(\ref{mstlog}) exists in all higher order terms of
1978: the $Z\alpha$-expansion. Ref.
1979: \cite{milstein_sushkov_terekhov_Log_all_orders} found the
1980: corresponding contribution of the higher-order terms numerically.
1981: This latest calculation, also shown in Fig.\ref{seven}, is even
1982: closer to our results given by Eqs.(\ref{ddd}) and
1983: (\ref{interp}). The deviation from (\ref{interp}) is below $ 9
1984: \% $ for a wide range of atomic charges. Ref.
1985: \cite{milstein_sushkov_terekhov_Log_all_orders} claims their
1986: error to be of the order of $\sim 5 \% $; our estimate for the
1987: accuracy of Eq.(\ref{ddd}), which limits the error at large $Z$,
1988: is a {\em few} percent, see Section \ref{equality}. Within these
1989: errors the results of the two groups completely agree. The fact
1990: that the remaining discrepancy increases for heavy atoms
1991: indicates probably that it is due to terms of the
1992: $Z\alpha$-expansion still unaccounted for in
1993: \cite{milstein_sushkov_terekhov_Log_all_orders}.
1994: Overall, Fig.\ref{seven} shows good agreement of the recent results
1995: obtained by the two groups using different approaches. The sharp
1996: contradiction that existed during the initial stages of this
1997: research (compare the dashed-dotted line with the thick-dotted
1998: line) makes the latest convergence even more satisfying and
1999: trustworthy .
2000:
2001: There is an interesting physical link between the chiral
2002: invariance identity of Eq.(\ref{ddd}) and the perturbation theory
2003: approach of Eqs.(\ref{197}),(\ref{mstlog}). Remember that in
2004: deriving Eq.(\ref{final}) in the previous section we focused our
2005: attention at one stage of the consideration at small distances $r
2006: \le Z\alpha/m$. The claim was that if this region is proved to
2007: comply with Eq.(\ref{final}), then the outer region $r \ge
2008: Z\alpha/m$ would inevitably comply with it as well, see
2009: discussion after Eq.(\ref{rr}). A similar philosophy lies behind
2010: the logarithmic approximation in Eq. (\ref{mstlog}). The
2011: integration that leads to the logarithmic function in this
2012: equation is saturated at small distances. The omitted constant is
2013: related to larger distances, but it is of lesser importance and
2014: can, in the simplest approximation, be neglected. Thus in both
2015: approaches the region of small distances proves to be the most
2016: important.
2017:
2018: This link between the chiral invariance identity and the
2019: perturbation theory shows that Eq.(\ref{ddd}) should be valid when
2020: one restricts consideration to the logarithmic approximation.
2021: This means that the term $\propto Z^2 \alpha^3 \ln mr_\mathrm{N}$
2022: should exist not only in the PNC problem, see Eq.(\ref{mstlog}),
2023: but in the FNS problem for the $s_{1/2}$ and $p_{1/2}$ partial
2024: waves as well. Eq.(\ref{ddd}) predicts a linear relation between
2025: the three coefficients of these three terms.
2026: Ref.\cite{milstein_sushkov_terekhov_02} demonstrated that the three
2027: coefficients mentioned are all equal, in compliance with
2028: Eq.(\ref{ddd}). This fact can be considered either as a
2029: consequence of Eq.(\ref{ddd}), or, alternatively, as its
2030: independent verification, being fruitful either way.
2031:
2032:
2033:
2034: The picture of reliability presented above based on the complete
2035: accord of all available data for the self-energy corrections in
2036: atoms has recently been put to test {\em again} by Milstein {\em
2037: et al}. They have claimed in Ref.
2038: \cite{milstein_sushkov_terekhov_controversial_Log} that for small
2039: values of the nuclear charge $Z\simeq 1$ the self-energy
2040: corrections for the energy shift induced by the FNS in the
2041: $p_{1/2}$ partial wave are positive and very large, being two
2042: orders of magnitude greater than anticipated previously for the
2043: hydrogen atom. For the $p_{3/2}$ state the proposed enhancement
2044: is even stronger, four orders of magnitude. This conclusion, if
2045: correct, could be related to the PNC problem for light elements
2046: through the equality of \cite{kf_prl_02} mentioned above that
2047: binds together the radiative corrections for the PNC and FNS
2048: problems. It needs, however, to be pointed out that a
2049: questionable approach is applied in
2050: \cite{milstein_sushkov_terekhov_controversial_Log}. It uses the
2051: {\em small}-momenta expansion for the vertex operator. When this
2052: technique is used for the short-range potential there arises a
2053: contradiction because the short-range potential needs {\em large}
2054: momenta. Attempts to use a similar line of arguments for
2055: calculations of radiative corrections have been made previously,
2056: but it was recognized that they strongly overestimate the effect,
2057: by orders of magnitude. A necessity to be very careful, to avoid
2058: using the small-momenta asymptotic dealing with the short-range
2059: potentials, was clearly stated long time ago by Lepage {\it et
2060: al} \cite{lepage_81}. One is forced to presume therefore that
2061: conclusions of \cite{milstein_sushkov_terekhov_controversial_Log}
2062: are not convincing (probably they should be even called
2063: preliminary, if one uses the latter term as in
2064: Ref.\cite{milstein_sushkov_01}). Fortunately, these results do
2065: not influence either the above mentioned good numerical agreement
2066: that was achieved for heavy atoms, or the analytical agreement
2067: between the two groups within the $Z\alpha $-expansion that
2068: defines the PNC in light elements.
2069:
2070:
2071: Discussions in this section are almost entirely devoted to
2072: analytical methods, the only exception was made for the numerical
2073: data available for corrections to the energy shifts due to the
2074: FNS. The main reason for this is that direct numerical
2075: calculations for the self-energy corrections for the PNC
2076: amplitude have not yet been reported. Presumably the
2077: difficulties in this approach are significant, though not being
2078: experts in this area we would not speculate on the details. We
2079: wish to mention, however, that developments in this direction are
2080: quite desirable due to several reasons. One of them, purely
2081: theoretical, has come into existence only recently. Numerical
2082: studies can determine independently the accuracy of the chiral
2083: identity (\ref{ddd}). Knowing the limitations of this relation in
2084: the PNC problem, one would be able to estimate the accuracy of
2085: other relations of this type that can be derived and used in
2086: similar problems in the future.
2087:
2088: Summarizing, we discussed above two different methods for
2089: calculation of the self-energy corrections, one related to the
2090: chiral invariance identity (\ref{ddd}), the other one based on
2091: the $Z\alpha$-expansion. They give close results proving that the
2092: self-energy corrections to the PNC amplitude are large and
2093: negative. Most importantly, this fact brings the experimental
2094: data for the $6s - 7s $ PNC amplitude in $^{133}$Cs into
2095: agreement with the standard model.
2096:
2097:
2098:
2099:
2100:
2101: \section{ comparison with experimental
2102: data }
2103:
2104: Fig. \ref{eight} summarizes the data for the QED radiative
2105: corrections presenting the vacuum polarization correction
2106: $\delta^{( \mathrm { VP }) }_{\mathrm {PNC}}$ calculated with the
2107: help of Eq.(\ref{w})-(\ref{w2}) and the self-energy (plus vertex)
2108: radiative correction $ \delta^{( \mathrm { e-line }) }_{\mathrm
2109: {PNC}}$ described by Eq.(\ref{interp}), as well as the total
2110: radiative correction \begin{equation} \label{totalQED} \delta^
2111: {( \mathrm { tot }) }_{\mathrm {PNC}} = \delta^{( \mathrm { VP })
2112: }_{\mathrm {PNC}} + \delta^{( \mathrm { e-line }) }_{\mathrm
2113: {PNC}}.
2114: \end{equation}
2115: We see that both the vacuum polarization and the self-energy
2116: corrections rise quickly with the nuclear charge, but their
2117: opposite signs make the total correction smoother. We find
2118: $\delta^{( \mathrm { tot }) }_{\mathrm {PNC}} =-0.54 \% $ for the
2119: Cs atom and $\delta^{( \mathrm { tot }) }_{\mathrm {PNC}}= -0.70 \%
2120: $ for Pb,Tl and Bi.
2121:
2122:
2123: Let us discuss the implications of these results, first for the
2124: $6s-7s$ PNC amplitude in $^{133}$Cs. The standard model value for
2125: the nuclear weak charge for Cs \cite{hagivara_02} is
2126:
2127:
2128: \begin{equation}\label{QW}
2129: Q_W (^{133}{\mathrm Cs}) = \,-73.09 \,\pm\,(0.03)~.
2130: \end{equation}
2131: Ref. \cite{dzuba_02} refined previous calculations of Ref.
2132: \cite{dzuba_89}, extracting from the experimental PNC amplitude of
2133: Ref. \cite{wood_97} the weak charge
2134:
2135: \begin{eqnarray}\label{72.45}
2136: && Q_W ^{(\mathrm{C+B+N}) }(^{133}{\mathrm Cs}) =
2137: % \,-72.18\pm(0.29)_{\mathrm {expt}}\pm(0.36)_{\mathrm{theor}}~,
2138: \\ \nonumber
2139: &&-72.45\pm(0.29)_{\mathrm {expt} }\pm(0.36)_{\mathrm {theor} },
2140: \end{eqnarray}
2141: with the theoretical error $0.5\%$. This value includes the
2142: correlation and the Breit corrections, as well as the neutron
2143: skin corrections (superscript C+B+N), but does not take into
2144: account the radiative corrections \footnote{In order to compare
2145: Eq.(\ref{72.45}) with results of Ref.\cite{dzuba_02} one needs
2146: to extract from -72.45 the vacuum polarization correction $0.4
2147: \% $. This results in $Q_W ( ^{133}{\mathrm Cs} ) = -72.16$ in
2148: agreement with Eq.(43) of \cite{dzuba_02}.}. The neutron
2149: nuclear radius is slightly larger than its proton radius. This
2150: provides a possibility for the electron to interact with neutrons
2151: in the region outside of the proton core, where the wave function
2152: is slightly smaller than inside the nucleus. The corresponding
2153: reduction of the PNC amplitude is called the neutron skin
2154: correction in PNC. According to estimates of Derevianko
2155: \cite{derevianko_02} the neutron skin correction is
2156: approximately $-0.2 \% $ for Cs. The calculations of
2157: Ref.\cite{dzuba_02} took this effect into account.
2158:
2159: Eq.(\ref{72.45}) is consistent with $Q_W(^{133}{\mathrm Cs}) =
2160: \,-72.21 \,\pm\,(0.28)_{\mathrm {expt}}\,\pm\, (0.34)_{\mathrm
2161: {theor}}$ that was adopted in \cite{johnson_01} by taking the
2162: average of the results of
2163: Refs.\cite{dzuba_89,blundell_90,kozlov_01}, and accepting the
2164: theoretical error $0.4\%$ proposed in \cite{bennett_wieman_99}.
2165: This value for the weak charge includes the vacuum polarization
2166: correction $0.4 \%$. Extracting it and taking into account $-0.2
2167: \% $ for the nuclear skin correction one obtains $-72.35$ that is
2168: close to Eq.(\ref{72.45}).
2169:
2170: The radiative corrections derived from results presented in Fig.
2171: \ref{eight} are $-0.54\pm(0.10) \%$; the error reflects the
2172: uncertainty of the self-energy radiative correction.
2173: Eq.(\ref{72.45}) combined with the QED radiative corrections gives
2174:
2175: \begin{eqnarray}\label{72.84}
2176: && Q_W^{\mathrm {Total}}(^{133}{\mathrm Cs}) =
2177: %\,-72.71 \pm(0.29)_{\mathrm expt}\pm(0.39)_{\mathrm theor}~,
2178: \\
2179: \nonumber
2180: && \,-72.84 \pm(0.29)_{\mathrm {expt}}\pm(0.36)_{\mathrm {theor}}~.
2181: \end{eqnarray}
2182: where Total = Correlations + Breit corrections + Neutron skin
2183: correction + QED radiative corrections. The agreement with the
2184: standard model given in Eq.(\ref{QW}) is good. It is so good, in
2185: fact, that we hasten to remind the reader a historical aspect of
2186: the problem. The theory has come up with Eq.(\ref{72.84}) after
2187: a turbulent period of research during which the experimental data
2188: were widely anticipated to be in contradiction with the standard
2189: model. This makes the found ``unexpected''agreement more
2190: objective.
2191:
2192: Let us consider now the case of the thallium atom that was
2193: studied experimentally in
2194: \cite{berkeley_79,berkeley_81,berkeley_85,oxford_91_Tl,oxford_95,
2195: seattle_95}, the calculations were performed in
2196: \cite{Tl_87,kozlov_pra_01}, that give close results for the
2197: many-electron correlation. The corrections to the result of
2198: \cite{Tl_87} should include contributions of the Breit
2199: interaction -0.98\% \cite{0.98dzuba}. As was mentioned, the QED
2200: radiative corrections for Tl give $-0.70 \% $. If we adopt
2201: the neutron skin correction $-0.2 \% $, then the total
2202: theoretical result reads
2203:
2204: \begin{equation}
2205: \label{qtl}
2206: \!\! R= \mathrm{Im }( E_{\mathrm{PNC}}/M1 )=
2207: -15.64 (-Q_W/N) \! \cdot \! 10 ^{-8}.
2208: \end{equation}
2209: Comparing with the measured value of Ref.\cite{seattle_93} that
2210: predicts $R=-14.68(17)\cdot 10^{-8}$, we obtain the weak charge for
2211: $^{205}$Tl
2212:
2213: \begin{equation}
2214: \label{qwtk}
2215: Q_W(^{205}\mathrm{Tl})=-116.4 \pm
2216: (1.3)_\mathrm{expt} \pm (3.4)_\mathrm{theor}
2217: \end{equation}
2218: in good agreement with the standard model prediction
2219: \cite{hagivara_02}
2220:
2221: \begin{equation} \label{Tl_sm} Q_W(^{205}\mathrm{Tl})= -116.7(1) ~.
2222: \end{equation}
2223:
2224:
2225:
2226: \section{conclusion}
2227: \label{conclusion}
2228:
2229: The QED corrections for parity nonconservation have recently
2230: emerged as an important ingredient in the analyses of modern
2231: experimental data. The improvement in the experimental accuracy,
2232: particularly for the $6s-7s$ PNC amplitude in $^{133}$Cs, and the
2233: progress of atomic structure calculations revealed a possible gap
2234: between the experiment data and atomic structure calculations on
2235: one side and the standard model on the other.
2236:
2237: The QED corrections, which embrace the Breit corrections and the
2238: radiative corrections, reconcile the experimental data with the
2239: standard model. The area has evolved and developed very rapidly.
2240: Over the past couple of years, even over a year, the situation
2241: has changed dramatically. A short time ago one could be
2242: seriously contemplating the possibility to modify the standard
2243: model to accommodate a possible deviation of atomic experimental
2244: data with the standard model. Presently it has become clear that
2245: modifications of the model are to be postponed, one simply needs
2246: to calculate everything accurately. {\it Everything} here is
2247: essential. Different parts of the QED corrections are all very
2248: large and have different signs. One must be certain that all
2249: parts are included and accounted for properly. That is why we
2250: spent some time and space above presenting and comparing all
2251: available results for the radiative corrections. The conclusion
2252: is that {\em all} data for the radiative corrections fit together
2253: well. For each possible correction several different methods of
2254: calculation were employed by different groups with the same final
2255: conclusions. The self-energy corrections, that proved to be by
2256: far the most difficult ones, are known presently with an
2257: uncertainty of better than 10 \%, which is sufficient for the
2258: present day experimental accuracy. One can contemplate
2259: a significant reduction of this error, if required.
2260:
2261: The calculations of the radiative corrections for PNC were based
2262: on vast expertise accumulated in other related
2263: research areas that include the Lamb shift, the hyperfine
2264: interaction, the nuclear finite-size correction. In turn, some
2265: methods of calculations that have recently evolved in relation to
2266: PNC may find applications in other areas. One of them is the chiral
2267: invariance identity Eq.(\ref{ddd}), that was originally applied
2268: to express radiative corrections to PNC in terms of the
2269: corrections to energy shifts due to the finite nuclear size.
2270: Similar identities can be derived for other related problems.
2271: One can also mention a simple convenient expression that gives the
2272: variation of the electron wave function at the atomic nucleus in
2273: terms of the two first moments of the perturbative potential
2274: Eq.(\ref{final}).
2275:
2276: In conclusion, large QED radiative corrections to the parity
2277: nonconservation amplitude in heavy atoms reconcile the
2278: accurate atomic experimental data on parity nonconservation for
2279: the $6s-7s$ PNC amplitude in $^{133}$Cs of Wood {\em et al}
2280: \cite{wood_97} with the standard model.
2281:
2282:
2283: % \acknoledgments
2284:
2285: Discussions with D.Budker, J.S.M.Ginges, W.Greiner, G.F.Gribakin,
2286: M.G.Kozlov, V.M.Shabaev, and A.V.Solov'yov are greatly
2287: appreciated. This work was supported by the Australian Research
2288: Council.
2289:
2290: %\bibliography
2291:
2292: \begin{thebibliography}{99}
2293:
2294:
2295: %%% ^{133}Cs 6s-7s
2296:
2297: \bibitem{bouchiat_74}
2298: % ^{133}Cs 6s-7s Paris
2299:
2300: M.A.Bouchiat and C.Bouchiat J.Phys. (Paris) {\bf 35}, 899 (1974);
2301: {\bf 36}, 493 (1974).
2302:
2303:
2304: \bibitem{bouchiat_82}
2305: % ^{133}Cs 6s-7s Paris
2306:
2307: M.A. Bouchiat, J. Gu{\'e}na, L. Hunter, and L. Pottier, Phys.
2308: Lett. B {\bf 117}, 358 (1982); errata {\bf 121}, 456 (1983).
2309:
2310:
2311: \bibitem{bouchiat_84}
2312: % ^{133}Cs 6s-7s Paris
2313:
2314: M.A. Bouchiat, J. Gu{\'e}na, L. Pottier, and L. Hunter, Phys.
2315: Lett. B {\bf 134}, 463 (1984).
2316:
2317: \bibitem{bouchiat_85}
2318: % ^{133}Cs 6s-7s Paris
2319:
2320: M.A. Bouchiat, J. Gu{\'e}na, and L. Pottier, J. Phys. (Paris) {\bf
2321: 46}, 1897 (1985).
2322:
2323: \bibitem{bouchiat_86}
2324: % ^{133}Cs 6s-7s Paris
2325:
2326: M.A. Bouchiat, J. Gu{\'e}na, and L. Pottier, J. Phys. (Paris)
2327: {\bf 47}, 1175 (1986).
2328:
2329:
2330: \bibitem{bouchiat_86_1}
2331: % ^{133}Cs 6s-7s Paris
2332:
2333: M.A. Bouchiat, J. Gu{\'e}na, L. Pottier, and L. Hunter, J.
2334: Phys. (Paris) {\bf 47}, 1709 (1986).
2335:
2336:
2337: \bibitem{bouchiat_02}
2338: % ^{133}Cs 6s-7s Paris
2339:
2340: J. Gu{\'e}na, D. Chauvat, Ph. Jacquier, E. Jahier, M. Lintz, A.V.
2341: Papoyan, S. Sanguinetti, D. Sarkisyan, A. Wasan, and M.A.
2342: Bouchiat, physics/0210069.
2343:
2344:
2345:
2346: \bibitem{boulder_85}
2347: %133 Cs 6s-7s Boulder
2348:
2349: S.L. Gilbert, M.C. Noecker, R.N. Watts, and C.E. Wieman, Phys.
2350: Rev. Lett. {\bf 55}, 2680 (1985).
2351:
2352:
2353: \bibitem{boulder_86}
2354: %133 Cs 6s-7s Boulder
2355:
2356: S.L. Gilbert and C.E. Wieman, Phys. Rev. A {\bf 34}, 792 (1986).
2357:
2358:
2359: \bibitem{boulder_88}
2360: %133 Cs 6s-7s Boulder
2361:
2362: M.C. Noecker, B.P. Masterson, and C.E. Wieman, Phys. Rev. Lett.
2363: {\bf 61}, 310 (1988).
2364:
2365: \bibitem{wood_97}
2366: %133 Cs 6s-7s Boulder
2367:
2368: C.S. Wood, S.C. Bennett, D. Cho, B.P. Masterson, J.L. Roberts,
2369: C.E. Tanner, and C.E. Wieman, Science {\bf 275}, 1759 (1997).
2370:
2371:
2372: %%% 205 Tl 6p,1/2 - 7p,1/2
2373:
2374: \bibitem{berkeley_79}
2375: % 205 Tl 6p,1/2 - 7p,1/2 Berkeley
2376:
2377: R. Conti, P. Bucksbaum, S. Chu, E. Commins, and L. Hunter, Phys.
2378: Rev. Lett. {\bf 42}, 343 (1979).
2379:
2380: \bibitem{berkeley_81}
2381: % 205 Tl 6p,1/2 - 7p,1/2 Berkeley
2382:
2383: P.K. Bucksbaum, E.D. Commins, and L.R. Hunter, Phys. Rev. Lett.
2384: {\bf 46}, 640 (1981);
2385:
2386: P.H. Bucksbaum, E.D. Commins, and L.R. Hunter, Phys. Rev. D {\bf
2387: 24}, 1134 (1981).
2388:
2389:
2390: \bibitem{berkeley_85}
2391: % 205 Tl 6p,1/2 - 7p,1/2 Berkeley
2392:
2393: P.S. Drell and E.D. Commins, Phys. Rev. Lett. {\bf 53}, 968
2394: (1984);
2395:
2396: P.S. Drell and E.D. Commins, Phys. Rev. A {\bf 32}, 2196 (1985).
2397:
2398:
2399:
2400:
2401: % Tl 6p_{1/2}-6p_{3/2}
2402:
2403: \bibitem{oxford_91_Tl}
2404: % Tl $6p_{1/2}-6p_{3/2}$ Oxford
2405:
2406: T.D. Wolfenden, P.E.G. Baird, and P.G.H. Sandars, Europhys. Lett.
2407: {\bf 15}, 731 (1991).
2408:
2409: \bibitem{oxford_95}
2410: % Tl 6p_{1/2}-6p_{3/2} Oxford
2411:
2412: N.H. Edwards, S.J. Phipp, P.E.G. Baird, and S. Nakayama, Phys.
2413: Rev. Lett. {\bf 74}, 2654 (1995).
2414:
2415:
2416: \bibitem{seattle_95}
2417: % Tl $6p_{1/2}-6p_{3/2}$ Seattle
2418:
2419: P.A. Vetter, D.M. Meekhof, P.K. Majumder, S.K. Lamoreaux, and E.N.
2420: Fortson, Phys. Rev. Lett. {\bf 74}, 2658 (1995).
2421:
2422:
2423:
2424: %%% Pb ^{3}P_{0}-^{3}P_{1}
2425:
2426:
2427: \bibitem{seattle_83}
2428: % Pb $^{3}P_{0}-^{3}P_{1}$ Seattle
2429:
2430: T.P. Emmons, J.M. Reeves, and E.N. Fortson, Phys. Rev. Lett. {\bf
2431: 51}, 2089 (1983); errata {\bf 52}, 86 (1984).
2432:
2433: \bibitem{seattle_93}
2434: % Pb $^{3}P_{0}-^{3}P_{1}$ Seattle
2435: D.M. Meekhof, P. Vetter, P.K. Majumder, S.K. Lamoreaux, and E.N.
2436: Fortson, Phys. Rev. Lett. {\bf 71}, 3442 (1993).
2437:
2438:
2439: \bibitem{oxford_96}
2440: % Pb $^{3}P_{0}-^{3}P_{1}$ Oxford
2441:
2442: S.J. Phipp, N.H. Edwards, P.E.G. Baird, and S. Nakayama, J. Phys.
2443: B {\bf 29}, 1861 (1996).
2444:
2445:
2446: %%% Bi $^{4}S_{3/2}-^{2}D_{5/2}$
2447:
2448:
2449: \bibitem{novosibirsk_78}
2450: % Bi $^{4}S_{3/2}-^{2}D_{5/2}$ Novosibirsk
2451:
2452: L.M. Barkov and M.S. Zolotorev, Pis'ma Zh. Eksp. Teor. Fiz. {\bf
2453: 27}, 379 (1978) [JETP Lett. {\bf 27}, 357 (1978)].
2454:
2455: \bibitem{novosibirsk_78_1}
2456: % Bi $^{4}S_{3/2}-^{2}D_{5/2}$ Novosibirsk
2457:
2458: L.M. Barkov and M.S. Zolotorev, Pis'ma Zh. Eksp. Teor. Fiz. {\bf
2459: 28}, 544 (1978) [JETP Lett. {\bf 28}, 503 (1978)].
2460:
2461: \bibitem{novosibirsk_79}
2462: % Bi $^{4}S_{3/2}-^{2}D_{5/2}$ Novosibirsk
2463:
2464: L.M. Barkov and M.S. Zolotorev, Phys. Lett. B {\bf 85}, 308
2465: (1979).
2466:
2467: \bibitem{novosibirsk_80}
2468: % Bi $^{4}S_{3/2}-^{2}D_{5/2}$ Novosibirsk
2469:
2470: L.M. Barkov and M.S. Zolotorev, Zh. Eksp. Teor. Fiz. {\bf 79}, 713
2471: (1980) [Sov. Phys. JETP {\bf 52}, 360 (1980)].
2472:
2473:
2474: \bibitem{moscow_84}
2475: % Bi $^{4}S_{3/2}-^{2}D_{5/2}$ Moscow
2476:
2477: G.N. Birich, Yu.V. Bogdanov, S.I. Kanorskii, I.I. Sobel'man, V.N.
2478: Sorokin, I.I. Struk, and E.A. Yukov, Zh. Eksp. Teor. Fiz. {\bf
2479: 87}, 776 (1984) [Sov. Phys. JETP {\bf 60}, 442 (1984)].
2480:
2481:
2482: \bibitem{oxford_87_5/2}
2483: % Bi $^{4}S_{3/2}-^{2}D_{5/2}$ Oxford
2484:
2485: J.D. Taylor, P.E.G. Baird, R.G. Hunt, M.J.D. Macpherson, G.
2486: Nowicki, P.G.H. Sandars, and D.N. Stacey, J. Phys. B {\bf 20},
2487: 5423 (1987).
2488:
2489:
2490: \bibitem{oxford_93}
2491: % Bi $^{4}S_{3/2}-^{2}D_{5/2}$ Oxford
2492:
2493: R.B. Warrington, C.D. Thompson, and D.N. Stacey, Europhys. Lett.
2494: {\bf 24}, 641 (1993).
2495:
2496:
2497:
2498: %%% Bi $^{4}S_{3/2}-^{2}D_{3/2}$
2499:
2500: \bibitem{seattle_81}
2501: % Bi $^{4}S_{3/2}-^{2}D_{3/2}$ Seattle
2502:
2503: J.H. Hollister, G.R. Apperson, L.L. Lewis, T.P. Emmons, T.G. Vold,
2504: and E.N. Fortson,
2505: Phys. Rev. Lett. {\bf 46}, 643 (1981).
2506:
2507:
2508:
2509: \bibitem{oxford_87_3/2}
2510: % Bi $^{4}S_{3/2}-^{2}D_{3/2}$ Oxford
2511:
2512: M.J.D. Macpherson, D.N. Stacey, P.E.G. Baird, J.P. Hoare, P.G.H.
2513: Sandars, K.M.J. Tregidgo, and Wang Guowen, Europhys. Lett. {\bf
2514: 4}, 811 (1987).
2515:
2516: \bibitem{oxford_91_Bi}
2517: % Bi $^{4}S_{3/2}-^{2}D_{3/2}$ Oxford
2518:
2519: M.J.D. Macpherson, K.P. Zetie, R.B. Warrington, D.N. Stacey, and
2520: J.P. Hoare, Phys. Rev. Lett. {\bf 67}, 2784 (1991).
2521:
2522: %%%%%%%%%%%%%%%%%%%
2523:
2524:
2525: \bibitem{khriplovich_91}
2526:
2527: I.B. Khriplovich. {\it Parity Nonconservation in Atomic
2528: pnenomena.} (Gordon and Breach, Philadelphia, 1991).
2529:
2530:
2531: \bibitem{dzuba_89}
2532:
2533: V.A.Dzuba, V.V.Flambaum, and O.P.Sushkov, Phys. Lett A {\bf 141},
2534: 147 (1989).
2535:
2536:
2537: \bibitem{blundell_90}
2538:
2539: S.A.Blundell, W.R.Johnson, and J.Sapirstein, Phys. Rev. Lett. {\bf
2540: 65}, 1411 (1990);
2541:
2542: S.A.Blundell, J.Sapirstein, and W.R.Johnson, Phys. Rev. D {\bf 45},
2543: 1602 (1992).
2544:
2545:
2546: \bibitem{bennett_wieman_99}
2547:
2548: S.C.Bennett and C.E.Wieman, Phys. Rev. Lett. {\bf 82}, 2484
2549: (1999); {\bf 82}, 4153 (1999); {\bf 83}, 889 (1999).
2550:
2551:
2552: \bibitem{kozlov_01}
2553:
2554: M.G.Kozlov, S.G.Porsev, and I.I.Tupitsyn, Phys. Rev. Lett. {\bf
2555: 86}, 3260 (2001).
2556:
2557:
2558: \bibitem{dzuba_01}
2559:
2560: V.A.Dzuba, V.V.Flambaum, and J.S.M. Ginges, hep-ph/0111019.
2561:
2562:
2563: \bibitem{dzuba_02}
2564:
2565: %High-precision calculation of parity nonconservation in cesium
2566: %and test of the Standard Model.
2567: V.A. Dzuba, V.V. Flambaum, J.S.M. Ginges. Phys. Rev. D66, 076013
2568: (2002); hep-ph/0204134.
2569:
2570: \bibitem{derevianko_00}
2571:
2572: A.Derevianko, Phys. Rev. Lett. {\bf 85}, 1618 (2000).
2573:
2574:
2575: \bibitem{dzuba_harabati_01}
2576:
2577: V.A.Dzuba, C.Harabati, W.R.Johnson, and M.S.Safronova, Phys. Rev A
2578: {\bf 63}, 044103 (2001).
2579:
2580:
2581: \bibitem{sushkov_01}
2582:
2583: O.P.Sushkov, Phys. Rev. A, {\bf 63}, 042504 (2001).
2584:
2585:
2586: \bibitem{johnson_01}
2587:
2588: W.R.Johnson, I.Bednyakov, and G.Soff, Phys. Rev. Lett. {\bf 87},
2589: 233001-1 (2001).
2590:
2591:
2592: \bibitem{milstein_sushkov_01}
2593:
2594: A.I.Milstein and O.P.Sushkov, Phys.Rev. A {\bf 66}, 022108/1-4
2595: (2002); hep-ph/0109257.
2596:
2597:
2598: \bibitem{kf_jpb_02}
2599:
2600: M.Yu.Kuchiev and V.V.Flambaum, J.Phys.B:At.Mol.Opt.Phys. {\bf 35},
2601: 4101 (2002); hep-ph/0205012.
2602:
2603:
2604: \bibitem{marciano_sirlin_83}
2605:
2606: W.J.Marciano and A.Sirlin, Phys. Rev. D {\bf 27}, 552 (1983).
2607:
2608:
2609: \bibitem{lynn_sandars_94}
2610:
2611: B.W.Lynn and P.G.H.Sandars, J. Phys. B {\bf 27}, 1469 (1994).
2612:
2613:
2614: \bibitem{blundell_97}
2615: S.A.Blundell, K.T.Cheng, and J.Sapirstein, Phys. Rev. A {\bf 55}, 1857
2616: (1997).
2617:
2618:
2619:
2620: \bibitem{sunnergren_98}
2621:
2622: P.Sunnergren, H.Persson, S.Salomonson, S.M.Sneider, I.Lindgren, and
2623: G.Soff, Phys. Rev. A {\bf 58}, 1055 (1998).
2624:
2625:
2626: \bibitem{kf_prl_02}
2627:
2628: M.Yu.Kuchiev and V.V.Flambaum, Phys.Rev.Lett. {\bf 89} 283002
2629: (2002); hep-ph/0206124.
2630:
2631:
2632: \bibitem{johnson_soff_85}
2633:
2634: W.R.Johnson and G.Soff, At. Data Nuc. Data Tables {\bf 33}, 405
2635: (1985).
2636:
2637:
2638: \bibitem{blundell_92}
2639:
2640: S.A.Blundell, Phys. Rev. A {\bf 46}, 3762 (1992).
2641:
2642:
2643: \bibitem{blundell_sapirstein_johnson_92}
2644:
2645: S.A.Blundell, J.Sapirstein, and W.R.Johnson, Phys. Rev. D {\bf 45},
2646: 1602 (1992).
2647:
2648:
2649: \bibitem{cheng_93}
2650:
2651: K.T.Cheng, W.R.Johnson and J.Sapirstein, Phys. Rev A {\bf 47}, 1817
2652: (1993).
2653:
2654:
2655: \bibitem{lindgren_93}
2656:
2657: I.Lindgren, H.Persson, S.Salomonson, and A.Ynnerman, Phys. Rev. A
2658: {\bf 47}, 4555 (1993).
2659:
2660:
2661: \bibitem{cheng_02}
2662:
2663: K.T.Cheng. Private communication (2002).
2664:
2665:
2666: \bibitem{k_jpb_02}
2667:
2668: M.Yu.Kuchiev, J.Phys.B:At.Mol.Opt.Phys. {\bf 35} 4101 (2002);
2669: hep-ph/0208196.
2670:
2671:
2672:
2673: \bibitem{milstein_sushkov_terekhov_02}
2674:
2675: A.I.Milstein, O.P.Sushkov, and I.S.Terekhov, Phys.Rev.Lett. {\bf
2676: 89} 28003 (2002); hep-ph/0208227.
2677:
2678:
2679:
2680: \bibitem{kf_comm_02}
2681:
2682: M.Yu.Kuchiev and V.V.Flambaum, hep-ph/0209052.
2683:
2684:
2685: \bibitem{milstein_sushkov_terekhov_Log_all_orders}
2686:
2687: A.I.Milstein, O.P.Sushkov, and I.S.Terekhov, hep-ph/0212072.
2688:
2689:
2690: \bibitem{milstein_sushkov_terekhov_controversial_Log}
2691:
2692: A.I.Milstein, O.P.Sushkov, and I.S.Terekhov, hep-ph/0212018
2693:
2694:
2695: \bibitem{lepage_81}
2696:
2697: G.P.Lepage, D.R.Yenni, and W.Erickson, Phys. Rev. Lett. {\bf 47},
2698: 1640 (1981).
2699:
2700:
2701: % \bibitem{expert}
2702:
2703: % We are forced to raise this issue in the present work long after
2704: % we have made our concerns known to the authors of
2705: % \cite{milstein_sushkov_terekhov_controversial_Log}. We hope that
2706: % attention to these concerns can help avoiding the unneccasary {\it
2707: % preliminary} results.
2708:
2709:
2710: \bibitem{LLIII}
2711: L.D. Landau and E.M. Lifshitz 1977
2712: {\it Quantum mechanics : non-relativistic theory}
2713: %translated from the Russian by J. B. Sykes and J. S. Bell
2714: (Oxford, New York, Pergamon Press)
2715:
2716:
2717:
2718: \bibitem{uehling}
2719:
2720: E.A.Uehling, Phys.Rev. {\bf 48}, 55 (1935).
2721:
2722:
2723:
2724: \bibitem{fullerton_rinker}
2725:
2726: L.W.Fullerton and J.A.Rinker, Jr., Phys.Rev.A {\bf 13}, 1283 (1976).
2727:
2728:
2729: \bibitem{wichmann-kroll}
2730:
2731: E.H.Wichmann and N.M.Kroll {\it Phys.Rev.} {\bf 101}, 843 (1956)
2732:
2733:
2734: \bibitem{milstein_strakhovenko_83}
2735:
2736:
2737: A.I.Milstein and V.M.Strakhovenko
2738: {\it Sov.ZhETF} {\bf 84}, 1247 (1983).
2739:
2740:
2741: \bibitem{df_03}
2742: V.A.Dzuba and V.V.Flambaum (to be published).
2743:
2744:
2745:
2746:
2747: \bibitem{pachucki_93}
2748:
2749: K.Pachucki, Phys. Rev. A {\bf 48}, 120 (1993).
2750:
2751:
2752: \bibitem{eides_97}
2753:
2754: M.I.Eides and H.Grotch, Phys. Rev. A {\bf 56}, R2507 (1997);
2755:
2756: M.I.Eides, H.Grotch, and V.A.Shelyuto, Phys. Rep. {\bf 342}, 63
2757: (2001).
2758:
2759:
2760: \bibitem{history}
2761:
2762: The work \cite{k_jpb_02} was inspired, not to say imposed on us,
2763: by a sceptical referee of \cite{kf_prl_02} who challenged us to
2764: support our results using perturbation theory.
2765: Ref.\cite{k_jpb_02} obliged dutifully, being written as a reply to
2766: the referee comment. The agreement between
2767: Refs.\cite{kf_prl_02,k_jpb_02} made our claim stronger, prompting
2768: us to acknowledge a fruitful, even if unnintentional, impact of our
2769: reluctant referee.
2770:
2771:
2772: \bibitem{sapirstein_03}
2773:
2774: J.Sapirstein, K.Pachucki, A.Vietia, and K.T.Cheng, hep-ph/0302202.
2775:
2776: \bibitem{hagivara_02}
2777: % D.E.Groom {\it et al}, Eur. Phys. J. C {\bf 15}, 1 (2000).
2778: K.Hagivara {\it et al}, Phys.Rev. D {\bf 64}, 010001 (2002).
2779:
2780:
2781: \bibitem{0.98dzuba}
2782:
2783: V.A.Dzuba, privet communication.
2784:
2785:
2786: \bibitem{derevianko_02}
2787: A.Derevianko, Phys.Rev. A {\bf 65}, 012106 (2002).
2788:
2789:
2790: \bibitem{Tl_87}
2791:
2792: V.A. Dzuba, V.V. Flambaum, P.G. Silvestrov, and O.P. Sushkov.
2793: J.Phys. B {\bf 20}, 3297 (1987).
2794:
2795:
2796: \bibitem{kozlov_pra_01}
2797:
2798: M.G. Kozlov, S.G. Porsev, and W.R. Johnson. Phys. Rev. A {\bf 64},
2799: 052107 (2001).
2800:
2801:
2802: \end{thebibliography}
2803:
2804: %%%%%%%%%%%%%%%%%%%%%%
2805:
2806:
2807:
2808:
2809:
2810: % \begin{center} {\bf Figure captions}
2811: % \end{center}
2812: % \vspace{1cm}
2813:
2814:
2815: \clearpage
2816: \pagebreak
2817: % \newpage
2818:
2819:
2820: \begin{figure}[t]
2821: \caption{ \label{one}
2822: %Fig.1.
2823: Relative contribution (\%) of the QED vacuum polarization to the
2824: PNC amplitude versus the nuclear charge. Thick line - total
2825: correction predicted by Eq.(\ref{w}), dashed line - prediction of
2826: Eq. (\ref{w1}) that takes into account only events outside the
2827: nucleus, dotted line - results based on the expansion
2828: Eq.(\ref{log2}).
2829: }\end{figure}
2830:
2831:
2832: % \vspace{2cm}
2833:
2834: \nopagebreak[4]
2835:
2836: \begin{figure}[h]
2837: \caption{ \label{two}
2838: %Fig.2.
2839: The Feynman diagrams for QED vertex (a) and self-energy (b),(c)
2840: corrections, called e-line corrections in the text. They are used
2841: in the text to describe either the PNC amplitude or the energy
2842: shifts induced by the finite nuclear size. The thick dot
2843: represents in these two cases either the vertex induced by the PNC
2844: Hamiltonian, or the finite nuclear size (FNS) potential that takes
2845: into account the spread of the nuclear charge inside the nucleus.
2846: The thick line describes the electron propagation in the atomic
2847: field.
2848: }\end{figure}
2849:
2850: % \vspace{1cm}
2851:
2852: \nopagebreak[4]
2853:
2854: \begin{figure}[h]
2855: \caption{\label{three}
2856: % Fig.3.
2857: The $s_{1/2}$ and $p_{1/2}$ wave functions for the valence
2858: electron in $^{133}$Cs (arbitrary units) versus the distance
2859: $r/r_\mathrm{C}$, where $r_\mathrm{C}=\hbar/mc$ is the Compton
2860: radius. The functions are normalized in such a way that the
2861: $const$ in both sets of boundary conditions in Eq.(\ref{bc}) takes
2862: the same value, $const = 1$. Solid and dotted lines - $f_{s,1/2}$
2863: and $g_{p,1/2}$ wave functions that are large for this
2864: normalization, thick dashed and thin dashed lines $-g_{s,1/2}$ and
2865: $f_{p,1/2}$ wave functions that are much smaller, (a) the region
2866: of large distances $r\sim r_\mathrm{C}$ shows good coincidence of
2867: $f_{s,1/2}$ with $g_{p,1/2}$, (b) the region in the vicinity of
2868: the nucleus, the nuclear radius is $r_\mathrm{N} \simeq
2869: 0.016~r_\mathrm{C} \simeq 6.1$ fm, where $-g_{s,1/2}$ coincides
2870: with $f_{p,1/2}$. The shown results illustrate Eq.(\ref{wf})
2871: based on chiral invariance. }\end{figure}
2872:
2873: \vspace{1cm}
2874:
2875: \pagebreak
2876:
2877: \begin{figure}[h]
2878: \caption{\label{four}
2879: %Fig.4.
2880: The relative radiative corrections ($\%$) induced by the e-line
2881: corrections (self-energy plus vertex) of Fig.\ref{two}. Thick,
2882: thin, and long-dashed lines - corrections for FNS energy shifts
2883: for $1s_{1/2}$, $2s_{1/2}$, and $2p_{1/2}$ levels respectively
2884: extracted from \cite{blundell_sapirstein_johnson_92}; short-dashed
2885: line - prediction of Eq.(\ref{ddd}) for the PNC matrix element
2886: calculated in \cite{kf_prl_02}, dashed-dotted line - correction to
2887: the hyperfine interaction \cite{kf_prl_02}.
2888: }\end{figure}
2889:
2890: % \vspace{1cm}
2891:
2892: \nopagebreak[4]
2893:
2894: \begin{figure}[h]
2895: \caption{ \label{five}
2896: %Fig.5
2897: The lowest order Feynman diagrams for the QED self-energy and
2898: vertex corrections, called e-line corrections in the text, to the
2899: PNC matrix element. The thin line describes propagation of a free
2900: electron. For each diagram one of the wavy legs shows the Coulomb
2901: interaction with the nucleus, another one - the weak PNC
2902: interaction with the nucleus (which was shown by a large dot in
2903: Fig. \ref{two} ). Each diagram represents all possible Feynman
2904: diagrams with the given topological structure.
2905: }\end{figure}
2906:
2907: % \vspace{1cm}
2908:
2909: %\pagebreak
2910:
2911: \begin{figure}[h]
2912: \caption{\label{six}
2913: %Fig.6
2914: Thick dotted, thin dotted and thick solid lines show e-line
2915: corrections (self-energy plus vertex) to the PNC amplitude
2916: predicted by Eq. (\ref{final}) of \cite{kf_prl_02}, by the linear
2917: approximation Eq.(\ref{197}) of \cite{k_jpb_02}, and by the
2918: interpolating Eq.(\ref{interp}) of \cite{k_jpb_02} (all same as in
2919: Fig. \ref{seven}). The thin dashed line, thick dashed line, and
2920: the dashed-dotted line - the e-line corrections to the FNS energy
2921: shifts in the linear approximation (\ref{FNS}), from results of
2922: \cite{blundell_sapirstein_johnson_92}, and the interpolating
2923: formula (\ref{intFNS}). Note the similarity between the PNC and
2924: FNS problems. }\end{figure}
2925:
2926: % \vspace{1cm}
2927:
2928:
2929:
2930: \begin{figure}[h] \caption{\label{seven}
2931: %Fig.7
2932: Relative contribution of the e-line corrections (self-energy
2933: plus vertex) to the PNC amplitude ( \% ) versus the nuclear
2934: charge. Thick dotted, thin dotted and thick solid lines show
2935: predictions of Eq. (\ref{final}) of \cite{kf_prl_02}, of the
2936: linear approximation (\ref{197}) of \cite{k_jpb_02}, and the
2937: interpolating Eq.(\ref{interp}) of \cite{k_jpb_02} respectively,
2938: as in Fig. \ref{six}. Dashed-dotted line, thin dashed line and
2939: thick dashed line - results of Milstein {\em et al} from
2940: \cite{milstein_sushkov_01,milstein_sushkov_terekhov_02,%
2941: milstein_sushkov_terekhov_Log_all_orders} respectively. Note
2942: the convergence of the latest results of the two groups, which
2943: strongly deviated at the initial stage. }\end{figure}
2944:
2945:
2946: % \vspace{1cm}
2947:
2948:
2949: \begin{figure}[h] \caption{\label{eight}
2950: %Fig.8
2951: QED radiative corrections to the PNC amplitude (\%) versus the
2952: nuclear charge. The dotted line - vacuum polarization Eq.(\ref{w}),
2953: thin line - self energy correction Eq.(\ref{interp}), thick line -
2954: total correction Eq.(\ref{totalQED}).
2955: }\end{figure}
2956:
2957:
2958:
2959:
2960:
2961:
2962: \end{document}
2963:
2964:
2965: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2966:
2967: