1: \documentclass[genTeX]{nrc1}
2:
3: \usepackage{cite}
4: \usepackage{amsmath}
5: \usepackage{amsfonts}
6: \usepackage{bm}
7: \usepackage{bbm}
8: \usepackage{epsf}
9: \usepackage{times}
10: %\usepackage{showkeys}
11: \usepackage{nicefrac}
12: \usepackage{graphicx}
13:
14: \def\buildrel#1\under#2{\mathrel{\mathop{\kern0pt #2}\limits_{#1}}}
15: \def\Res#1{{\buildrel {\scriptstyle #1} \under {\textstyle \rm Res}}\,}
16: \def\bbox#1{\mbox{\boldmath$#1$}}
17: \def\bm#1{\mbox{\boldmath$#1$}}
18: \def\eq{\;\;\Longleftrightarrow\;\;}
19: \def\corresponds{\sim}
20: \def\succsim{\succ\kern-.9em_\sim\kern.3em}
21: \def\precsim{\prec\kern-1em_\sim\kern.3em}
22: \def\slantfrac#1#2{\kern1em^{#1}\kern-.3em/\kern-.1em_{#2}}
23: \def\lfrac#1#2{{}^{#1\!}\kern-.0em/_{#2}}
24: \def\trunc#1{[\mkern - 2.5 mu [#1] \mkern - 2.5 mu ]}
25:
26: \sloppy{}
27:
28: \journal{Canadian Journal of Physics}
29:
30: \volyear{1}[2004]{2004}
31:
32: \bibliographystyle{myprsty}
33:
34: \setcounter{page}{1}
35:
36: \begin{document}
37:
38: \title{Some Recent Advances in Bound--State Quantum Electrodynamics}
39:
40: \author{U.~D.~Jentschura and J.~Evers}
41:
42: \address{Max--Planck--Institut f\"ur Kernphysik,
43: Saupfercheckweg 1, 69117 Heidelberg, Germany}
44:
45: \maketitle
46:
47: \shortauthor{Jentschura and Evers}
48:
49: \begin{abstract}
50: We discuss recent progress in various problems
51: related to bound-state quantum electrodynamics:
52: the bound-electron $g$ factor,
53: two-loop self-energy corrections
54: and the laser-dressed Lamb shift.
55: The progress relies on various advances
56: in the bound-state formalism, including
57: ideas inspired by effective field theories
58: such as Nonrelativistic Quantum Electrodynamics.
59: Radiative corrections in dynamical processes
60: represent a promising field for further
61: investigations.
62: \end{abstract}
63:
64: \keywords{Calculations and mathematical techniques in atomic and
65: molecular physics, \\
66: quantum electrodynamics -- specific calculations;\\
67: {\it PACS numbers:} 31.15.-p, 12.20.Ds}
68:
69: %
70: % Introduction
71: %
72: \section{Introduction}
73:
74: This is a brief summary of a number of recent advances
75: in our understanding of bound-state quantum electrodynamic
76: (QED) effects. The topics are (i) two-loop corrections
77: to the bound-electron $g$ factor, (ii) higher-order
78: two-loop corrections to the self-energy of a bound electron,
79: and (iii) the laser-dressed Lamb shift.
80:
81: The first two of these rather diverse topics are
82: related to two-loop effects. The investigation
83: of these is simplified considerably by the use of
84: effective field-theory techniques inspired
85: by Nonrelativistic QED (NRQED)~\cite{CaLe1986,Pa1993,PiSo1998}.
86: The Wilson coefficients multiplying the effective
87: operators in the NRQED Lagrangian are matched against
88: those of the full relativistic theory, providing a simplified
89: framework for the calculation of bound-state effects.
90: Scale-separation parameters such as the photon mass $\mu$
91: are cancelled at the end of the calculation.
92: The analysis of higher-order corrections to the $g$ factor
93: of the bound electron is simplified further by
94: a transformation to the length gauge, which results in
95: a lesser number of terms to be considered than would be
96: necessary in the velocity gauge. This fact has inspired
97: the development of Long-wavelength QED (LWQED)~\cite{Pa2004},
98: a theory which is obtained after Power--Zienau and
99: Foldy--Wouthuysen transformations of the first-quantized
100: Lagrangian; the second quantization is carried out by formulating
101: the path integral. Consequently, an improved understanding
102: and a tremendous simplification results for the
103: calculation of a number of QED corrections for bound states,
104: such as the $g$ factor and higher-order corrections to the
105: self-energy.
106:
107: A further field of recent studies has been concerned with
108: the interaction of a laser-dressed bound electron with the radiation
109: field~\cite{JeEvHaKe2003,JeKe2004aop,EvJeKe2004,JeEvKe2004}.
110: This process entails corrections which can only be understood
111: if the analysis is carried out right from the start
112: in the framework of the laser-dressed states, which are
113: the eigenstates of the quantized atom-laser Hamiltonian
114: in the rotating-wave approximation~\cite{CT1975}.
115:
116: %
117: % Bound-electron $g$ factor
118: %
119: \section{Bound-electron $g$ factor}
120:
121: In this section, we briefly summarize the
122: results of a recent investigation~\cite{PaJeYe2004}
123: of the bound-electron $g$ factor, which is based on
124: NRQED. The central result of this investigation is the
125: following semianalytic expansion
126: in powers of $Z\alpha$ and $\ln(Z\alpha)$
127: for the bound-electron $g$ factor ($n{\rm S}$ state)
128: in the non-recoil limit, which is
129: the limit of an infinite nuclear mass:
130: %
131: \begin{align}
132: \label{gbound}
133: g(n{\rm S}) & =
134: \underbrace{2 - \frac{2\, (Z\alpha)^2}{3\,n^2} +
135: \frac{(Z\alpha)^4}{n^3} \, \left( \frac{1}{2 n} - \frac23 \right) +
136: {\cal O}{(Z\alpha)^6}}_{\mbox{Breit (1928),
137: Dirac theory}}
138: \nonumber\\[4ex]
139: &
140: + \underbrace{{\frac{\alpha}{\pi}}\,
141: \left\{ {2 \times \frac{1}{2}}\, {\bigg(}
142: 1 + \frac{(Z\alpha)^2}{6 n^2} {\bigg)}
143: + \frac{(Z\alpha)^4}{n^3} {\bigg\{}
144: a_{41}\, {\ln}[(Z\alpha)^{-2}] +
145: a_{40} {\bigg\}}
146: + {\cal O}(Z\alpha)^5 \right\}}_{\mbox{one-loop correction}}
147: \nonumber\\[4ex]
148: &
149: + \underbrace{{ \left(\frac{\alpha}{\pi}\right)^2}\,
150: \left\{ {-0.656958} \, {\bigg(} 1 +
151: \frac{(Z\alpha)^2}{6 n^2} {\bigg)} +
152: \frac{(Z\alpha)^4}{n^3} {\bigg\{}
153: b_{41}\,
154: {\ln}[(Z\alpha)^{-2}] +
155: b_{40} {\bigg\}}
156: + {\cal O}(Z\alpha)^5 \right\}}_{\mbox{two-loop correction}}
157: \nonumber\\[4ex]
158: & + {\cal O}(\alpha^3)\,.
159: \end{align}
160: %
161: This expansion is valid through the order
162: of two loops (terms of order $\alpha^3$ are
163: neglected). The notation is in part inspired by the usual conventions
164: for Lamb-shift coefficients: the (lower case) $a$ terms denote the
165: one-loop effects, with $a_{kj}$ denoting the
166: coefficient of a term proportional to
167: $\alpha\,(Z\alpha)^k\,{\ln}^j[(Z\alpha)^{-2}]$.
168: The $b$ terms denote the two-loop corrections,
169: with $b_{kj}$ multiplying
170: a term proportional to
171: $\alpha^2\,(Z\alpha)^k\,{\ln}^j[(Z\alpha)^{-2}]$.
172: In~\cite{PaJeYe2004}, complete results are derived for the
173: coefficients $a_{41}$, $a_{40}$ and $b_{41}$.
174:
175: In general, the expression corresponding to (\ref{gbound})
176: for a free electron is obtained by letting the
177: parameter $Z\alpha \to 0$ in every term of the loop
178: expansion (expansion in powers of $\alpha$).
179: In this limit, the known free-electron two-loop result is
180: recovered~\cite{KaKr1950,So1957,So1958,Pe1957helv,Re1972a,Re1972b,Ad1989}.
181:
182: \begin{figure}
183: \begin{center}
184: \begin{minipage}{10cm}
185: \begin{center}
186: \includegraphics[width=0.8\linewidth]{fig1.eps}
187: \end{center}
188: \caption{\label{fig1} Feynman diagrams
189: for the two-loop self-energy corrections
190: to the bound-electron $g$-factor.}
191: \end{minipage}
192: \end{center}
193: \end{figure}
194:
195: Up to the relative order $(Z\alpha)^2$, the free-electron
196: contribution in one-, two-, and higher-loop
197: order is multiplied by a relative factor
198: %
199: \begin{equation}
200: \label{Zalpha2}
201: 1 + (Z\alpha)^2 \, a_{20} =
202: 1 + (Z\alpha)^2 \, b_{20} = 1 + \frac{(Z\alpha)^2}{6 n^2}\,.
203: \end{equation}
204: %
205: This result consequently holds for the three-loop and the
206: four-loop term not shown in Eq.~(\ref{gbound}).
207: The applicability of the relative factor (\ref{Zalpha2})
208: to the two-loop term, valid through $(Z\alpha)^2$,
209: had been stressed previously in~\cite{Ka2001proc}.
210: The result in Eq.~(\ref{Zalpha2}) had been obtained originally
211: in~\cite{Gr1970prl,Gr1970pra,GrHe1971,Fa1970plb,Fa1970nc}
212: (for the 1S state).
213: As is evident from Eq.~(\ref{gbound}),
214: the correction of relative order $(Z\alpha)^2$
215: is different on the level of the tree-level diagrams
216: and reads $g(n{\rm S}) \sim 2 \times [1 - (Z\alpha)^2/(3 n^2)]$.
217:
218: Explicit results for the coefficients in (\ref{gbound}),
219: restricted to the one-loop self-energy, read~\cite{PaJeYe2004}
220: %
221: \begin{subequations}
222: \label{a4}
223: \begin{align}
224: \label{a41}
225: a_{41}(n{\rm S}) &= \frac{32}{9}\,, \\[2ex]
226: \label{a40}
227: a_{40}(n{\rm S}) &= \frac{73}{54}
228: - \frac{5}{24 n}
229: - \frac{8}{9} \, \ln k_0(n{\rm S})
230: - \frac{8}{3} \, \ln k_3(n{\rm S})\,.
231: \end{align}
232: \end{subequations}
233: %
234: Here, $\ln k_0(nS)$ is the Bethe logarithm for an $nS$ state,
235: and $\ln k_3(nS)$ is a generalization of the Bethe logarithm
236: to a perturbative potential of the form $1/r^3$.
237: Vacuum polarization adds a further $n$-independent
238: contribution of $(-16/15)$ to $a_{40}$~\cite{Ka2000}.
239: The Bethe logarithms for 1S and 2S~\cite{DrSw1990}
240: read
241: %
242: \begin{subequations}
243: \label{lnk0}
244: \begin{eqnarray}
245: \label{lnk01S}
246: \ln k_0(1{S}) &=& 2.984~128~555\,,\\[1ex]
247: \label{lnk02S}
248: \ln k_0(2{S}) &=& 2.811~769~893\,,
249: \end{eqnarray}
250: \end{subequations}
251: %
252: and the corresponding values for $\ln k_3$ read~\cite{PaJeYe2004}
253: %
254: \begin{subequations}
255: \label{lnk3}
256: \begin{eqnarray}
257: \label{lnk31S}
258: \ln k_3(1{S}) &=& 3.272~806~545\,,\\[1ex]
259: \label{lnk32S}
260: \ln k_3(2{S}) &=& 3.546~018~666\,.
261: \end{eqnarray}
262: \end{subequations}
263: %
264: The quantity $\ln k_3(n{\rm S})$ is defined as,
265: %
266: \begin{equation}
267: \int_0^\epsilon {\rm d} k\, k^2\,
268: \langle\phi|\bm{r}\,\frac{1}{E-H_S-k}\,\frac{1}{r^3}\,
269: \frac{1}{E-H_S-k}\,\bm{r}\,|\phi\rangle
270: = -4\,\frac{(Z\alpha)^3}{n^3} \,
271: \left[\ln \frac{\mu}{(Z\alpha)^2}+\frac56-\ln k_3\right]
272: \end{equation}
273: %
274: where the ultraviolet cutoff $\epsilon$ is to be understood
275: in the sense of~\cite{Pa1993}, and the matching of the noncovariant
276: cutoff $\epsilon$ to the covariant photon mass $\mu$
277: is given as (see~\cite{ItZu1980}, pp.~361--362)
278: %
279: \begin{equation}
280: \ln \left( \frac{2\epsilon}{m} \right) \to
281: \ln \left(\frac{\mu}{m}\right) + \frac{5}{6}\,.
282: \end{equation}
283: %
284: However, this replacement is not unique and the constant
285: term depends on the actual form of the integrand. A different
286: replacement has to be used for some of the low-energy photon
287: corrections to the $g$ factor~\cite{PaJeYe2004}.
288:
289: The results for the two-loop coefficients read
290: %
291: \begin{subequations}
292: \label{b4}
293: \begin{align}
294: \label{b41}
295: b_{41}(n{\rm S}) &= \frac{56}{9}\,, \\[2ex]
296: \label{b40}
297: b_{40}(1{\rm S}) &= -18.5(5.5)\,.
298: \end{align}
299: \end{subequations}
300: %
301: Here, the result for $b_{40}$ is an estimate based
302: on an explicit calculation of a large contribution due to
303: low-energy virtual photons, and an estimate
304: of the remaining, unknown contribution due to high-energy virtual
305: photons. The dominant logarithmic two-loop
306: term $b_{41}$ is caused
307: exclusively by the two-loop self-energy diagrams in Fig.~\ref{fig1}
308: alone. The other two-loop diagrams,
309: which include closed fermion loops,
310: can be found in Fig.~21 of~\cite{Be2000}.
311: The logarithmic term $b_{41}$ is, however,
312: exclusively related to the gauge-invariant
313: subset displayed in Fig.~\ref{fig1}.
314:
315: The newly calculated $a_{41}$, $a_{40}$ and $b_{41}$
316: are bound-state corrections to the electron $g$-factor
317: of order $(\alpha/\pi)\, (Z\alpha)^4$ and
318: $(\alpha/\pi)^2\, (Z\alpha)^4$, multiplied by
319: logarithmic terms. These corrections
320: are (at $Z=1$) formally of order $\alpha^5$ and $\alpha^6$
321: and therefore of the same order of magnitude as
322: the tenth- and twelfth-order corrections to the
323: free-electron anomaly, which barely are of experimental
324: or theoretical significance at the current level
325: of accuracy. One may therefore ask why these binding
326: corrections are of any phenomenological significance.
327: The reason is that at somewhat higher $Z$,
328: the situation changes drastically, due to $Z^4$
329: scaling of the binding corrections. In addition,
330: due to numerically large coefficients and logarithmic
331: factors, the ``hierarchy'' of the corrections changes drastically.
332: Roughly, one may say that at $Z=1$, the bound-electron
333: anomalous magnetic moment is approximately independent
334: of binding corrections of order $(\alpha/\pi)\, (Z\alpha)^4$
335: and higher, whereas for higher $Z$, the situation is reversed,
336: and the binding corrections to the one- and two-loop
337: contributions are numerically much more
338: significant than the higher-loop
339: free-electron corrections. This ``transition
340: from free to bound-state quantum electrodynamics'' as a
341: function of $Z$ is a somewhat
342: peculiar feature of the bound-electron $g$-factor.
343:
344: For, example, we consider the ratio
345: %
346: \begin{equation}
347: \label{r1def}
348: r_1(Z) = \frac{\displaystyle
349: \left(\frac{\alpha}{\pi}\right)\,(Z\alpha)^4\,\ln[(Z\alpha)^{-2}]}
350: {\displaystyle \left(\alpha/\pi\right)^4}\,,
351: \end{equation}
352: %
353: which gives an order-of-magnitude estimate for the the ratio of the
354: one-loop self-energy binding correction to the
355: eighth-order anomalous magnetic moment of the free electron.
356: We have
357: %
358: \begin{equation}
359: \label{r1num}
360: r_1(Z=1) \approx 2\,, \qquad
361: r_1(Z=10) \approx 10^4 \gg 1\,. \qquad
362: \end{equation}
363: %
364: For the two-loop logarithmic binding correction, we
365: have
366: %
367: \begin{equation}
368: \label{r2def}
369: r_2(Z) = \frac{\displaystyle
370: \left(\frac{\alpha}{\pi}\right)^2\,(Z\alpha)^4\,\ln[(Z\alpha)^{-2}]}
371: {\displaystyle \left(\alpha/\pi\right)^4}
372: \end{equation}
373: %
374: and consequently
375: %
376: \begin{equation}
377: \label{r2num}
378: r_2(Z=1) \approx 0.2 < 1\,, \qquad
379: r_2(Z=10) \approx 2 \times 10^3 \gg 1 \,. \qquad
380: \end{equation}
381: %
382: As is evident from these considerations, the
383: $(Z\alpha)^4$ one-loop and two-loop binding corrections are
384: roughly of the same order of magnitude as the
385: highly problematic four-loop
386: corrections~\cite{KiLi1990,Ki1995ieee,HuKi1999} for the free
387: electron. However, the situation changes drastically even at
388: very moderate nuclear charge numbers, and the binding
389: corrections to the one-loop and two-loop contributions
390: become dominant over the higher-loop effects.
391:
392: The NRQED one-loop calculation~\cite{PaJeYe2004} is divided into
393: three parts, the first of which entails fully relativistic
394: form-factor corrections including lower-order terms,
395: the second of which corresponds to a spin-dependent
396: scattering amplitude, and the third of which is a
397: low-energy Bethe-logarithm type correction that contains
398: $\ln k_0$ and $\ln k_3$. The one-loop
399: correction in Eq.~(\ref{gbound}) can therefore be
400: written in a natural way as
401: $ \delta g^{(1)} = g_1^{(1)} + g_2^{(1)} + g_3^{(1)}$,
402: where
403:
404: \begin{subequations}
405: \label{g1}
406: \begin{align}
407: \label{g11}
408: g_1^{(1)} &= \frac{\alpha}{\pi}\biggl[ 1+\frac{(Z\alpha)^2}{6\,n^2} -
409: \frac{(Z\alpha)^4}{n^3}\,\left(\frac76
410: +\frac5{24n} +\frac{16}{3}\, \ln \mu \right) \biggr] \,, \\[2ex]
411: %
412: \label{g21}
413: g_2^{(1)} &= \frac{\alpha}{\pi}\,\frac{(Z\alpha)^4}{n^3}\,
414: \Bigl(4+\frac{16}{9}\,\ln\mu\Bigr)\,, \\[2ex]
415: %
416: \label{g31}
417: g_3^{(1)} &= \frac{\alpha}{\pi}\,\frac{(Z\alpha)^4}{n^3}\,\frac{32}{9}\,
418: \biggl[\ln\frac{\mu}{(Z\alpha)^2} -\frac{5}{12} -\frac{\ln k_0}{4}
419: -\frac{3}{4}\,\ln k_3 \biggr]\,.
420: \end{align}
421: \end{subequations}
422:
423: The new contribution of order $(Z\alpha)^4$ can be
424: compared with the numerical results for the self-energy
425: correction~\cite{YeInSh2002,YeInSh2004} complete to all orders in $Z\alpha$.
426: Assuming correctness of the logarithmic term in Eq.~(\ref{g1}),
427: a fit to numerical data yields
428: $a_{40}^{(1)}({\rm 1S}) = -10.2(1)$ and
429: $a_{40}^{(1)}({\rm 2S}) = -10.6(1.2)$ for the constant term, in excellent
430: agreement with the analytic results which read
431: $a_{40}^{(1)}({\rm 1S}) = -10.236~524~318(1)$
432: and $a_{40}^{(1)}({\rm 2S}) = -10.707~715~607(1)$ .
433:
434: Having verified the consistency
435: of the analytic~\cite{PaJeYe2004}
436: and numerical results~\cite{YeInSh2002,YeInSh2004}, an interpolation
437: procedure~\cite{IvKa2001proc} may now be used to extract
438: a more accurate theoretical prediction
439: at low and intermediate nuclear charge numbers,
440: if combined with numerical results at
441: higher $Z$~\cite{YeInSh2002}. Thus, the
442: results in Eqs.~(\ref{a4}) and (\ref{b4})
443: may be used in order to infer improved theoretical
444: predictions for the bound-electron $g$ factor,
445: notably in the experimentally important
446: special cases of hydrogenlike carbon~\cite{HaEtAl2000prl}
447: and oxygen~\cite{VeEtAl2004}. Alternatively, the
448: improved status of the theory may be used in order to
449: infer a more accurate value of the
450: electron mass. Specifically, the value from
451: the carbon measurement~\cite{HaEtAl2000prl},
452: using the new theory, reads
453: %
454: \begin{equation}
455: m(^{12}{\rm C}^{5+}) = 0.000\,548\,579\,909\,41\, (29)(3)\ {\rm u} \,.
456: \end{equation}
457: %
458: The first error comes from the
459: experiment~\cite{HaEtAl2000prl}, and the
460: second error corresponds to the theoretical uncertainty.
461: The conclusion is that a further
462: improvement of the experiment could lead to a much
463: better determination of the electron
464: mass; the new theory provides room
465: for at least an improved determination by one order of magnitude.
466:
467: For the calculation of yet higher-order binding corrections
468: to the one-loop and two-loop contributions,
469: a detailed understanding of the two-loop form-factors,
470: including their slopes, is required.
471: The most recent calculations of these effects,
472: in both dimensional and photon-mass regularizations,
473: can be found in~\cite{BoMaRe2003,MaRe2003,BoMaRe2004}.
474:
475: \begin{figure}[htb]
476: \begin{center}
477: \begin{minipage}{14cm}
478: \begin{center}
479: \includegraphics[width=1.0\linewidth]{fig2.eps}
480: \end{center}
481: \caption{\label{fig2} The two-loop corrections to the
482: Lamb shift in hydrogenlike systems fall naturally
483: into three gauge-invariant subsets, which are the
484: pure two-loop self-energy terms (2LSE), the
485: vacuum-polarization correction to the virtual-photon
486: line in the self-energy (SVPE), and the
487: self-energy vacuum-polarization and pure
488: two-loop vacuum-polarization corrections (SEVP).
489: The two-loop Bethe logarithm is a numerically
490: large correction to the nonlogarithmic term of order
491: $\alpha^2 (Z\alpha)^6$ and it is exclusively related
492: to the 2LSE subset; however, for a complete result
493: in this order, contributions from the other gauge-invariant subsets
494: will have to be considered as well.}
495: \end{minipage}
496: \end{center}
497: \end{figure}
498:
499: %
500: % Two-loop corrections
501: %
502: \section{Two-loop Bethe logarithms}
503:
504: As is well known~\cite{ApBr1970,Ka1993log,Pa1994prl,%
505: EiSh1995,Ka1996,Ka1996b,Ka1997,Pa2001},
506: the two-loop Lamb shift $\Delta E^{(2)}$,
507: in the limit of an infinite nuclear mass, may be written as
508: %
509: \begin{eqnarray}
510: \Delta E^{(2)} =
511: \left(\frac{\alpha}{\pi}\right)^2 \,
512: \frac{(Z\alpha)^4 \, m_{\rm e} \,c^2}{n^3} \,
513: H(Z\alpha)\,.
514: \end{eqnarray}
515: %
516: For S states, the dimensionless function $H(Z\alpha)$,
517: has a semianalytic expansion of the form
518: %
519: \begin{align}
520: H(Z\alpha) = & B_{40} + (Z\alpha) \, B_{50}
521: \nonumber\\[2ex]
522: & + (Z\alpha)^2 \, \left[ B_{63} \,
523: \ln^3[(Z\alpha)^{-2}] + B_{62} \, \ln^2[(Z\alpha)^{-2}]
524: + B_{61} \, \ln[(Z\alpha)^{-2}] + B_{60} \right] \,,
525: \end{align}
526: %
527: where we ignore higher-order terms, and upper case is used
528: for the $B_{ij}$ coefficients that multiply terms
529: of order $\alpha^2\,(Z\alpha)^i\,\ln^j[(Z\alpha)^{-2}] \, m_{\rm e} \,c^2$.
530: The coefficients, restricted to the
531: two-photon self-energy diagrams (Fig.~\ref{fig2}),
532: read as follows
533: %
534: \begin{subequations}
535: \begin{eqnarray}
536: B^{\rm (2LSE)}_{63}(n{\rm S}) &=&
537: -\frac{8}{27} = -0.296296\,, \\[2ex]
538: B^{\rm (2LSE)}_{62}(1{\rm S}) &=&
539: \frac{16}{27} - \frac{16}{9}\, \ln(2)
540: = -0.639\,669\,, \\[2ex]
541: %
542: B^{\rm (2LSE)}_{62}(n{\rm S}) &=&
543: B^{\rm (2LSE)}_{62}(1{\rm S}) +
544: \frac{16}{9} \left( \frac34 + \frac{1}{4 n^2} -
545: \frac1n - \ln(n) \! + \! \Psi(n) + C\right)\,.
546: \end{eqnarray}
547: \end{subequations}
548:
549: The $n$-dependence of $B_{61}$ has been
550: clarified in~\cite{Pa2001,Je2003jpa}
551: %
552: \begin{eqnarray}
553: B^{\rm (2LSE)}_{61}(n{\rm S}) &=& B^{\rm (2LSE)}_{61}(1{\rm S})
554: + \frac43 \, \left[ N(n{\rm S}) - N(1{\rm S})\right]
555: \nonumber\\[0.5ex]
556: & & + \! \left( \frac{80}{27} - \frac{32}{9} \, \ln (2)\right)
557: \left(\frac34 + \frac{1}{4 n^2} - \frac1n - \ln(n) \! + \! \Psi(n)
558: \! + \! C\right)\,,
559: \end{eqnarray}
560: %
561: where $C=0.577216\dots$ is Euler's constant,
562: $\Psi(n)$ is the logarithmic derivative of the Gamma function,
563: and $N(n{\rm S})$ is related to a correction to the
564: Bethe logarithm induced by a Dirac-delta potential.
565: Explicit values for $N(n{\rm S})$ can be found in~\cite{Je2003jpa}
566: ($n = 1,\dots,8$).
567:
568: \begin{figure}[htb]
569: \begin{center}
570: \begin{minipage}{8cm}
571: \begin{center}
572: \includegraphics[width=0.9\linewidth]{fig3.eps}
573: %
574: \caption{\label{fig3}
575: The integration regions for the two virtual photons in the
576: two-loop self-energy problem comprise a low-energy regime
577: with two low-energy photons, which gives rise to $b_L$.
578: The middle-energy regions with one low-energy and a one
579: high-energy photon give rise to $b_M$. The
580: high-energy contribution $b_F + b_H$ is as yet unknown.}
581: \end{center}
582: \end{minipage}
583: \end{center}
584: \end{figure}
585:
586: The $B_{60}$ coefficients are the sum of several
587: contributions
588: %
589: \begin{subequations}
590: \begin{eqnarray}
591: \label{2lse}
592: B^{\rm (2LSE)}_{60}(n{\rm S}) &=&
593: b_L + b_M + \left\{ b_F + b_H + b_{\rm VP} \right\}\,, \\[2ex]
594: b_L &=& b_L(n{\rm S}) \; \corresponds \;
595: \mbox{two-loop Bethe logarithm, two soft photons}\,,
596: \\[2ex]
597: b_M &=& b_M(n{\rm S}) \; = \;
598: \frac{10}{9} \, N(n{\rm S})
599: \; \corresponds \; \mbox{one soft, one hard photon}\,,
600: \\[2ex]
601: b_F &\corresponds&
602: \mbox{soft electron momenta, two hard photons}\,,
603: \\[2ex]
604: b_H &\corresponds&
605: \mbox{hard electron momenta, two hard photons}\,,
606: \\[2ex]
607: b_{\rm VP} &\corresponds&
608: \mbox{vacuum-polarization corrections}\,.
609: \end{eqnarray}
610: \end{subequations}
611: %
612: Only the terms $b_L + b_M$ are currently
613: known~\cite{PaJe2003,Je2004b60}~(see also Fig.~\ref{fig2}).
614: The contributions in curly brackets in
615: Eq.~(\ref{2lse}) remain to be evaluated.
616: However, an estimate for the total value of $B_{60}$
617: may be obtained,
618: %
619: \begin{eqnarray}
620: B_{60}(n{\rm S}) &=&
621: b_L + b_M + \left\{ b_F + b_H + b_{\rm VP} \right\}\,, \nonumber\\[2ex]
622: B_{60}(n{\rm S}) &\approx&
623: b_L + b_M \; \pm \; 15\,\%\,.
624: \end{eqnarray}
625: %
626: This estimate~\cite{PaJe2003,Je2004b60} is based on corresponding
627: one-loop calculations,
628: where the low-energy virtual photons give the by far
629: dominant contribution to the constant term~\cite{Pa1993}.
630: The results for the two-loop Bethe logarithms of S states
631: read~\cite{PaJe2003,Je2004b60}
632: %
633: \begin{subequations}
634: \label{bL}
635: \begin{eqnarray}
636: b_L(1S) &=& -81.4(3)\,, \\[2ex]
637: b_L(2S) &=& -66.6(3)\,, \\[2ex]
638: b_L(3S) &=& -63.5(6)\,, \\[2ex]
639: b_L(4S) &=& -61.8(8)\,, \\[2ex]
640: b_L(5S) &=& -60.6(8)\,, \\[2ex]
641: b_L(6S) &=& -59.8(8)\,.
642: \end{eqnarray}
643: \end{subequations}
644: %
645: A few clarifying remark might be in order. The $B_{60}$ coefficient
646: multiplies a correction of order
647: $\alpha^2 (Z\alpha)^6$, which is effectively an order-$\alpha^8$
648: contribution to the energy levels of hydrogen ($Z=1$).
649: In order to complete the calculation at this
650: order of magnitude, it would also be necessary to
651: consider the four-loop Dirac form-factor slope of the electron,
652: as well as the three-loop binding correction
653: of order $\alpha^3\,(Z\alpha)^5\,m_{\rm e}\,c^2$~\cite{EiSh2004}.
654: The three-loop slope has recently been evaluated
655: in~\cite{MeRi2000}, completing the theory of
656: energy levels in hydrogen up to the order of $\alpha^7$.
657:
658: %
659: % Laser-dressed Lamb shift
660: %
661: \section{Laser-dressed Lamb shift}
662:
663: In the recent past, seminal advances have been obtained both
664: in the techniques
665: of high-precision spectroscopy (e.g.,~\cite{ReEtAl2000}),
666: and in the coherent preparation
667: and manipulation of media by external electromagnetic
668: fields~\cite{ScZu1997,FiSw2005}.
669: Thus it is desirable to study the bound electrons
670: interacting simultaneously both with the quantized
671: radiation field and with an external driving
672: field. An accurate theory of such systems,
673: including all dynamic effects, might eventually
674: open a possibility for a whole new class of high-precision
675: experiments, provided that technical problems related
676: to the required highly accurate intensity stabilization of the laser
677: (and others) can be solved.
678: Traditionally, radiative and relativistic corrections
679: are treated with methods of QED, whereas studies
680: related to the dynamical nature of the interaction of matter
681: with driving laser fields are the domain of Quantum Optics (QO).
682: Obviously, a treatment of bound electrons in the presence
683: of both the radiation field and external driving fields requires
684: a combination of ideas from both subject areas:
685: While, {\em a priori}, the essential-state approximation
686: of QO~\cite{ScZu1997} is not sufficient to obtain the accuracy
687: of QED, a perturbative treatment of the interaction of the bound
688: electron with a strong external (laser) field as in QED is hopeless
689: because of the large coupling parameter.
690:
691: In~\cite{JeEvHaKe2003,JeKe2004aop,EvJeKe2004},
692: an atom with two relevant energy levels driven by a strong
693: near-resonant monochromatic laser field is studied as the easiest
694: model system for the above problem. The incoherent
695: part of the resonance fluorescence spectrum emitted by this system
696: in QO is known as the Mollow spectrum, where the coupling strength
697: is characterized by the the Rabi frequency
698: $\Omega$ defined as ($\hbar = \epsilon_0 = c = 1$)
699: %
700: \begin{equation}
701: \Omega = -q \left <e|{\bm x}\cdot {\bm \epsilon}_{\rm L}|g \right >
702: {\cal E}_{\rm L}\,,
703: \end{equation}
704: %
705: for a driving laser field ${\bm E}_{\rm L}(t) = {\cal E}_{\rm L} {\bm
706: \epsilon}_{\rm L} \cos (\omega_{\rm L}t)$ with frequency $\omega_{\rm L}$,
707: macroscopic classical amplitude ${\cal E}_{\rm L}$ and
708: polarization ${\bm \epsilon}_{\rm L}$. Here,
709: $q=-|q|$ is the elementary charge.
710: The corresponding coupling constant $g_{\rm L}$ for
711: the interaction of a quantized driving laser field with the
712: main atomic transition
713: is defined by
714: %
715: \begin{equation}
716: g_{\rm L} =
717: - q \, \langle g | \bm{\epsilon}_{\rm L} \cdot \bm{x} | e \rangle \,
718: {\cal E}_{\rm L}^{\rm (\gamma)}\, ,
719: \end{equation}
720: %
721: where ${\cal E}_{\rm L}^{\rm (\gamma)} =
722: \sqrt{\omega_{\rm L}/2V}$ is the electric laser field per photon
723: and $V$ is the quantization volume. The matching of the electric
724: field per photon with the corresponding classical macroscopic
725: electric field ${\cal E}_{\rm L}$ is then given by
726: %
727: \begin{equation}
728: \label{matching}
729: 2 \: \sqrt{n+1} \: {\cal E}_{\rm L}^{\rm (\gamma)} \:
730: \longleftrightarrow \: {\cal E}_{\rm L} \,.
731: \end{equation}
732: %
733:
734: If $\Omega$
735: is larger than the natural decay width $\Gamma$ of the
736: transition, then the Mollow spectrum approximately consists of
737: one central and two sideband peaks of Lorentzian shape, which
738: are located symmetrically around the driving laser field frequency.
739: The sideband peaks are shifted from the driving field frequency
740: by the generalized Rabi frequency
741: $\Omega_{\textrm R} = \sqrt{\Omega^2 + \Delta^2}$,
742: where $\Delta=\omega_{\textrm L} - \omega_{\textrm R}$
743: is the detuning of the laser field frequency $\omega_{\textrm L}$
744: with regard to the atomic transition frequency $\omega_{\textrm R}$.
745: The shape of the Mollow spectrum
746: may easily be explained in terms of the dressed states, which
747: are defined as the eigenstates of the quantum optical interaction
748: picture Hamiltonian describing the matter-light interaction.
749: In transferring to the dressed state picture, the interaction with
750: the driving laser field is accounted for to all orders.
751:
752: Thus, when evaluating radiative and relativistic corrections
753: to the Mollow spectrum, it is natural to start the analysis
754: from the dressed-state basis as opposed to the
755: unperturbed atomic bare-state basis.
756: In~\cite{JeEvHaKe2003,JeKe2004aop}, it was shown
757: that this distinction in fact has to be made. It is not
758: sufficient to modify the energies (which enter in the
759: formula for the dressed states) according to the usual
760: bare-state Lamb shift in order to obtain the correct
761: result for the corrections to the Mollow spectrum. Instead,
762: at nonvanishing detuning and nonvanishing Rabi frequency,
763: a treatment starting from the dressed-state basis leads
764: to an additional nontrivial correction term. This term
765: gives rise to a shift of the Mollow sidebands relative
766: to the central peak given by
767: %
768: \begin{equation}
769: \label{deltaomegaplusminus}
770: \delta \omega_\pm^{(C)} =
771: \mp {\cal C}\, \frac{\Omega^2}{\sqrt{\Omega^2 + \Delta^2}}\,,
772: \end{equation}
773: %
774: where
775: %
776: \begin{eqnarray}
777: {\cal C} &=& \frac{\alpha}{\pi} \, \ln[(Z\alpha)^{-2}] \,
778: \frac{\left< \bm{p}^2 \right>_g + \left< \bm{p}^2 \right>_e}{m^2}
779: \end{eqnarray}
780: %
781: is a dimensionless constant. Here, the notation $\langle . \rangle_g$
782: and $\langle . \rangle_e$ denotes the expectation value evaluated
783: with the ground or excited atomic state, respectively.
784:
785: Inspired by the interpretation of the bare Lamb shift correction
786: in terms of a ``summed'' shift as
787: in~\cite{JeEvHaKe2003,JeKe2004aop}, this additional correction
788: can be interpreted as a modification
789: to the Rabi frequency:
790: %
791: \begin{equation}
792: \label{summedC}
793: \delta \overline{\omega}^{(C)}_\pm = \pm
794: \left( \sqrt{\Omega^2 \, (1 - {\cal C})^2 +
795: \Delta ^2} - \sqrt{\Omega^2 + \Delta^2}\right)\,,
796: \end{equation}
797: %
798: with $\delta \overline{\omega}^{(C)}_\pm \approx
799: \delta\omega_\pm^{(C)}$ because of the
800: smallness of the correction. It should be noted that this
801: interpretation in terms of a summation is not trivial
802: and was shown to be valid up to first order in the correction.
803:
804: In~\cite{JeEvHaKe2003,EvJeKe2004}, the leading relativistic
805: and radiative corrections up to relative orders $(Z\alpha)^2$
806: and $\alpha(Z\alpha)^2$, respectively, have been evaluated,
807: as well as all other relevant correction terms up to the
808: specified order of approximation. It turns out that all corrections
809: may be interpreted as either corrections to the Rabi frequency
810: $\Omega$ or as corrections to the detuning $\Delta$, such that one
811: can define the fully corrected generalized Rabi frequency
812: $\Omega^{(j)}_{\mathcal C}$ by
813: %
814: \begin{equation}
815: \label{fullOmega}
816: \Omega^{(j)}_{\mathcal C} =
817: \sqrt{\Omega^2 \cdot \left(1+{\hat{\Omega}_{\rm rad}^{(j)}}\right)^2 +
818: \left(\Delta - {\Delta^{(j)}_{\rm rad}}\right)^2}\,.
819: \end{equation}
820: %
821: Here, $\hat{\Omega}_{\rm rad}^{(j)}$ contains
822: all corrections to the Rabi frequency, namely
823: the relativistic and radiative corrections to the transition dipole
824: moment, field-configuration dependent corrections, higher-order
825: corrections to the self-energy, and corrections to the secular
826: approximation.
827: $\Delta^{(j)}_{\rm rad}$
828: consists of all corrections to the detuning, i.e.
829: the bare Lamb shift, Bloch-Siegert shifts, and off-resonant
830: radiative corrections.
831: The superscript $(j)$ indicates the dependence
832: of the result on the total angular momentum quantum number.
833:
834: Equation~(\ref{fullOmega}) summarizes the main result of
835: this study: In the presence of driving laser fields, the usual
836: bare state Lamb shift of the atomic states is augmented
837: by additional correction terms. These in part depend on the
838: laser field parameters $\Omega$ and $\Delta$, which span
839: a two-dimensional parameter manifold determining the
840: actual value of the dynamical Lamb shift.
841:
842: A promising candidate for the experiment are
843: the hydrogen $1S_{\nicefrac{1}{2}}\leftrightarrow 2P_{\nicefrac{1}{2}}$ and
844: $1S_{\nicefrac{1}{2}}\leftrightarrow 2P_{\nicefrac{3}{2}}$
845: transitions.
846: We consider here
847: as a specific example the
848: $1S_{\nicefrac{1}{2}}\leftrightarrow 2P_{\nicefrac{1}{2}}$ transition
849: with $\Omega = 1000\cdot \Gamma_{\nicefrac{1}{2}}$ and
850: $\Delta = 50 \cdot \Gamma_{\nicefrac{1}{2}}$ as the
851: laser field parameters. The Rabi frequency is shifted
852: with respect to the leading-order
853: expression $\Omega_{\rm R} = \sqrt{\Omega^2 + \Delta^2}$ by
854: relativistic and radiative corrections as follows,
855: %
856: \begin{equation}
857: \pm \left ( \Omega^{(\nicefrac{1}{2})}_{\mathcal C} - \Omega_{\rm R} \right )
858: = \pm 738.282(60)\cdot 10^6 \: \rm{Hz}\, .
859: \end{equation}
860: %
861: This driving laser field parameter set is expected to be within reach
862: of improvements of the currently available Lyman-$\alpha$ laser
863: sources~\cite{EiWaHa2001} in the next few years.
864: The corresponding result for the
865: $1S_{\nicefrac{1}{2}}\leftrightarrow 2P_{\nicefrac{3}{2}}$ transition
866: with $\Omega = 1000\cdot \Gamma_{\nicefrac{3}{2}}$, $\Delta = 50 \cdot \Gamma_{\nicefrac{3}{2}}$
867: is
868: \begin{equation}
869: \pm \left ( \Omega^{(\nicefrac{3}{2})}_{\mathcal C} - \Omega_{\rm R} \right )
870: = \pm 734.871(60)\cdot 10^6 \: \rm{Hz}\, .
871: \end{equation}
872: %
873: All given uncertainties are due to unknown higher-order
874: terms~\cite{EvJeKe2004}.
875:
876: By a comparison to experimental data,
877: one may verify the presence of dynamical leading-logarithmic correction
878: to the dressed-state radiative shift in Eq.~(\ref{deltaomegaplusminus}),
879: which cannot
880: be explained in terms of the bare Lamb shift alone. This allows to
881: address
882: questions related to the physical reality of the dressed states.
883: On the other hand, the comparison with experimental results could
884: also be used
885: to interpret the nature of the evaluated radiative corrections in the sense
886: of the summation formulas which lead to the interpretation of
887: the shifts as arising from relativistic and radiative corrections
888: to the detuning and the Rabi frequency.
889:
890: %
891: % Laser-dressed Lamb shift
892: %
893: \section{Conclusions}
894:
895: One of the obvious conclusions to
896: be drawn from the recent advances in bound-state
897: quantum electrodynamics is as follows.
898: A widespread opinion has been invalidated
899: which suggested that two-loop corrections
900: to the bound-electron $g$ factor, and two-loop
901: corrections to the Lamb shift in higher order
902: might be a severe, if not insurmountable,
903: obstacle against further theoretical progress.
904: Quite to the contrary, the recent advances have shown
905: that an understanding of these effects is feasible
906: to a good accuracy. While some important
907: contributions remain to be evaluated,
908: the further program (e.g., the evaluation of
909: the remaining high-energy parts)
910: is clearly defined, and decisive first
911: steps toward a much improved understanding
912: have been accomplished.
913:
914: Further advances are possible both on the
915: experimental as well as on the theoretical side:
916: concerning the $g$ factor, the recent
917: theoretical progress allows for an order-of-magnitude improvement
918: of the value for the electron mass based on a potential new measurement alone.
919: Regarding the Lamb shift in hydrogen and low-$Z$
920: hydrogenlike systems, one has recently gained
921: an improved understanding of the binding corrections of order
922: $\alpha^2 \,(Z\alpha)^6\, \ln^j[(Z\alpha)^{-2}]\,
923: m_{\rm e}\,c^2$ (where $j$ may assume the values $j = 0,\dots,3$).
924: Recent numerical investigations in the regime
925: of intermediate nuclear charge numbers~\cite{YeInSh2004nim}
926: have also contributed toward an improvement of our understanding.
927: The extrapolation of the low-$Z$ results
928: by deferred Pad\'{e} approximants~\cite{Je2003plb} suggests
929: a rather good general consistency of both approaches,
930: while some issues regarding the consistency of the
931: analytic and numerical approaches remain to be
932: addressed~\cite{YeInSh2004nim}.
933:
934: Other progress concerns the modifications of radiative
935: corrections in dynamical processed as opposed to
936: $S$-matrix energy shifts. The laser-dressed
937: Lamb shift~\cite{JeEvHaKe2003,JeKe2004aop,EvJeKe2004,JeEvKe2004}
938: is a dynamical correction to the dressed-state~\cite{CT1975}
939: quasi-energies. The self-energy, in this case,
940: gives the by far dominant effect. While the bulk of the
941: laser-dressed Lamb shift can be understood in terms
942: of a radiative correction to the detuning, which is
943: taken into account in a natural way by evaluating the
944: detuning in terms of the observed (low-intensity)
945: resonance frequency of the transition, a few tiny
946: shifts persist which can only be understood if the
947: system is treated in the dressed-state picture right
948: from the start. The dynamic Lamb shift is an
949: effect which depends on two parameters that determine
950: the dynamics of the system: (i) the detuning and (ii) the
951: Rabi frequency. Therefore,
952: the dynamical Lamb shift could be mapped out as a function
953: of these parameters in a possible experiment.
954: A rather promising candidate for a possible measurement
955: would be based on the hydrogen 1S--2P$_{\nicefrac{1}{2}}$
956: transition~\cite{JeEvHaKe2003}. However, a very attractive
957: alternative would be provided by a forbidden M1 transition~\cite{DrEtAl2003}
958: in Ar XIV, $2s^2 \,2p\,
959: {}^2{\rm P}_{\nicefrac{1}{2}}$--${}^2{\rm P}_{\nicefrac{3}{2}}$,
960: provided the many-body QED effects can be treated to sufficient
961: accuracy. The system in question is described very well by
962: a two-level formalism, and the resonance line width
963: is small.
964:
965: \section*{Acknowledgments}
966:
967: The authors acknowledge insightful discussions with
968: C. H. Keitel, V. A. Yerokhin and K. Pachucki.
969: Helpful conversations with S. G. Karshenboim are also
970: gratefully acknowledged.
971:
972: \begin{thebibliography}{10}
973:
974: \bibitem{CaLe1986}
975: W.~E. Caswell and G.~P. Lepage, Phys. Lett. B {\bf 167}, 437 (1986).
976:
977: \bibitem{Pa1993}
978: K. Pachucki, Ann. Phys. (N.Y.) {\bf 226}, 1 (1993).
979:
980: \bibitem{PiSo1998}
981: A. Pineda and J. Soto, Phys. Rev. D {\bf 59}, 016005 (1998).
982:
983: \bibitem{Pa2004}
984: K. Pachucki, Phys. Rev. A {\bf 69}, 052502 (2004).
985:
986: \bibitem{JeEvHaKe2003}
987: U.~D. Jentschura, J. Evers, M. Haas, and C.~H. Keitel, Phys. Rev. Lett. {\bf
988: 91}, 253601 (2003).
989:
990: \bibitem{JeKe2004aop}
991: U.~D. Jentschura and C.~H. Keitel, Ann. Phys. (N.Y.) {\bf 310}, 1 (2004).
992:
993: \bibitem{EvJeKe2004}
994: J. Evers, U.~D. Jentschura, and C.~H. Keitel, Relativistic and Radiative
995: Corrections to the Mollow Spectrum, e-print quant-ph/0403202, submitted to
996: Phys. Rev. A.
997:
998: \bibitem{JeEvKe2004}
999: U.~D. Jentschura, J. Evers, and C.~H. Keitel, Relativistic and Radiative
1000: Corrections to Multi-Level Mollow--Type Spectra, Laser Physics, at press
1001: (special issue on the occasion of the 70${}^{\rm th}$ birthday of Prof.
1002: Herbert Walther).
1003:
1004: \bibitem{CT1975}
1005: C. Cohen-Tannoudji, in {\em Aux fronti\`{e}res de la spectroscopie
1006: laser/Frontiers in Laser Spectroscopy}, edited by R. Balian, S. Haroche, and
1007: S. Liberman (North-Holland, Amsterdam, 1975), pp.\ 4--104.
1008:
1009: \bibitem{PaJeYe2004}
1010: K. Pachucki, U.~D. Jentschura, and V.~A. Yerokhin, Phys. Rev. Lett. {\bf 93},
1011: 150401 (2004).
1012:
1013: \bibitem{KaKr1950}
1014: R. Karplus and N.~M. Kroll, Phys. Rev. {\bf 77}, 536 (1950).
1015:
1016: \bibitem{So1957}
1017: C.~M. Sommerfield, Phys. Rev. {\bf 107}, 328 (1957).
1018:
1019: \bibitem{So1958}
1020: C.~M. Sommerfield, Ann. Phys. (N.Y.) {\bf 5}, 26 (1958).
1021:
1022: \bibitem{Pe1957helv}
1023: A. Petermann, Helv. Phys. Acta {\bf 30}, 407 (1957).
1024:
1025: \bibitem{Re1972a}
1026: E. Remiddi, Nuovo Cim. A {\bf 11}, 825 (1972).
1027:
1028: \bibitem{Re1972b}
1029: E. Remiddi, Nuovo Cim. A {\bf 11}, 865 (1972).
1030:
1031: \bibitem{Ad1989}
1032: G.~S. Adkins, Phys. Rev. D {\bf 39}, 3798 (1989).
1033:
1034: \bibitem{Ka2001proc}
1035: S.~G. Karshenboim, in {\em The Hydrogen Atom}, edited by S.~G. Karshenboim and
1036: F.~S. Pavone (Springer, Berlin, 2001), pp.\ 651--663.
1037:
1038: \bibitem{Gr1970prl}
1039: H. Grotch, Phys. Rev. Lett. {\bf 24}, 39 (1970).
1040:
1041: \bibitem{Gr1970pra}
1042: H. Grotch, Phys. Rev. A {\bf 2}, 1605 (1970).
1043:
1044: \bibitem{GrHe1971}
1045: H. Grotch and R.~A. Hegstrom, Phys. Rev. A {\bf 4}, 59 (1971).
1046:
1047: \bibitem{Fa1970plb}
1048: R. Faustov, Phys. Lett. B {\bf 33}, 422 (1970).
1049:
1050: \bibitem{Fa1970nc}
1051: R. Faustov, Nuovo Cim. {\bf 69}, 37 (1970).
1052:
1053: \bibitem{Ka2000}
1054: S.~G. Karshenboim, Phys. Lett. A {\bf 266}, 380 (2000).
1055:
1056: \bibitem{DrSw1990}
1057: G.~W.~F. Drake and R.~A. Swainson, Phys. Rev. A {\bf 41}, 1243 (1990).
1058:
1059: \bibitem{ItZu1980}
1060: C. Itzykson and J.~B. Zuber, {\em Quantum Field Theory} (McGraw-Hill, New York,
1061: NY, 1980).
1062:
1063: \bibitem{Be2000}
1064: T. Beier, Phys. Rep. {\bf 339}, 79 (2000).
1065:
1066: \bibitem{KiLi1990}
1067: T. Kinoshita and W.~B. Lindquist, Phys. Rev. D {\bf 42}, 636 (1990).
1068:
1069: \bibitem{Ki1995ieee}
1070: T. Kinoshita, IEEE Trans. Instrum. Meas. {\bf 44}, 498 (1995).
1071:
1072: \bibitem{HuKi1999}
1073: V.~W. Hughes and T. Kinoshita, Rev. Mod. Phys. {\bf 71}, S133 (1999).
1074:
1075: \bibitem{YeInSh2002}
1076: V.~A. Yerokhin, P. Indelicato, and V.~M. Shabaev, Phys. Rev. Lett. {\bf 89},
1077: 143001 (2002).
1078:
1079: \bibitem{YeInSh2004}
1080: V.~A. Yerokhin, P. Indelicato, and V.~M. Shabaev, Phys. Rev. A {\bf 69},
1081: 052503 (2004).
1082:
1083: \bibitem{IvKa2001proc}
1084: V.~G. Ivanov and S.~G. Karshenboim, in {\em The Hydrogen Atom}, edited by
1085: S.~G. Karshenboim and F.~S. Pavone (Springer, Berlin, 2001), pp.\ 637--650.
1086:
1087: \bibitem{HaEtAl2000prl}
1088: H. H\"{a}ffner, T. Beier, N. Hermanspahn, H.-J. Kluge, W. Quint, J. Verd\'{u},
1089: and G. Werth, Phys. Rev. Lett. {\bf 85}, 5308 (2000).
1090:
1091: \bibitem{VeEtAl2004}
1092: J. Verd\'{u}, S. Djeki\'{c}, S. Stahl, T. Valenzuela, M. Vogel, G. Werth, T.
1093: Beier, H.~J. Kluge, and W. Quint, Phys. Rev. Lett. {\bf 92}, 093002 (2004).
1094:
1095: \bibitem{BoMaRe2003}
1096: R. Bonciani, P. Mastrolia, and E. Remiddi, Nucl. Phys. B {\bf 661}, 289
1097: (2003).
1098:
1099: \bibitem{MaRe2003}
1100: P. Mastrolia and E. Remiddi, Nucl. Phys. B {\bf 664}, 341 (2003).
1101:
1102: \bibitem{BoMaRe2004}
1103: R. Bonciani, P. Mastrolia, and E. Remiddi, Nucl. Phys. B {\bf 676}, 399
1104: (2004).
1105:
1106: \bibitem{ApBr1970}
1107: T. Appelquist and S.~J. Brodsky, Phys. Rev. A {\bf 2}, 2293 (1970).
1108:
1109: \bibitem{Ka1993log}
1110: S.~G. Karshenboim, JETP {\bf 76}, 541 (1993), [ZhETF {\bf 103}, 1105 (1993)].
1111:
1112: \bibitem{Pa1994prl}
1113: K. Pachucki, Phys. Rev. Lett. {\bf 72}, 3154 (1994).
1114:
1115: \bibitem{EiSh1995}
1116: M.~I. Eides and V.~A. Shelyuto, Phys. Rev. A {\bf 52}, 954 (1995).
1117:
1118: \bibitem{Ka1996}
1119: S.~G. Karshenboim, J. Phys. B {\bf 29}, L29 (1996).
1120:
1121: \bibitem{Ka1996b}
1122: S.~G. Karshenboim, JETP {\bf 82}, 403 (1996), [ZhETF {\bf 109}, 752 (1996)].
1123:
1124: \bibitem{Ka1997}
1125: S.~G. Karshenboim, Z. Phys. D {\bf 39}, 109 (1997).
1126:
1127: \bibitem{Pa2001}
1128: K. Pachucki, Phys. Rev. A {\bf 63}, 042503 (2001).
1129:
1130: \bibitem{Je2003jpa}
1131: U.~D. Jentschura, J. Phys. A {\bf 36}, L229 (2003).
1132:
1133: \bibitem{PaJe2003}
1134: K. Pachucki and U.~D. Jentschura, Phys. Rev. Lett. {\bf 91}, 113005 (2003).
1135:
1136: \bibitem{Je2004b60}
1137: U.~D. Jentschura, Two-Loop Bethe Logarithms for Higher Excited $S$ Levels,
1138: Phys. Rev. A (2004), at press.
1139:
1140: \bibitem{EiSh2004}
1141: M.~I. Eides and V.~A. Shelyuto, Phys. Rev. A {\bf 70}, 022506 (2004).
1142:
1143: \bibitem{MeRi2000}
1144: K. Melnikov and T. v.~Ritbergen, Phys. Rev. Lett. {\bf 84}, 1673 (2000).
1145:
1146: \bibitem{ReEtAl2000}
1147: J. Reichert, M. Niering, R. Holzwarth, M. Weitz, T. Udem, and T.~W. H\"{a}nsch,
1148: Phys. Rev. Lett. {\bf 84}, 3232 (2000).
1149:
1150: \bibitem{ScZu1997}
1151: M.~O. Scully and M.~S. Zubairy, {\em Quantum Optics} (Cambridge University
1152: Press, Cambridge, 1997).
1153:
1154: \bibitem{FiSw2005}
1155: Z. Ficek and S. Swain, {\em Quantum Interference and Coherence --- Springer
1156: Series in Optical Sciences vol. 100} (Springer, Berlin, Heidelberg, New York,
1157: to appear in 2005, ISBN: 0-387-22965-5).
1158:
1159: \bibitem{EiWaHa2001}
1160: K.~S.~E. Eikema, J. Walz, and T.~W. H\"{a}nsch, Phys. Rev. Lett. {\bf 86},
1161: 5679 (2001).
1162:
1163: \bibitem{YeInSh2004nim}
1164: V.~A. Yerokhin, P. Indelicato, and V.~M. Shabaev, Two-loop self-energy
1165: correction to the ground-state Lamb shift in H-like ions, e-print
1166: hep-ph/0409048.
1167:
1168: \bibitem{Je2003plb}
1169: U.~D. Jentschura, Phys. Lett. B {\bf 564}, 225 (2003).
1170:
1171: \bibitem{DrEtAl2003}
1172: I. Dragani\'{c}, J.~R. Crespo L\'{o}pez-Urrutia, R. DuBois, S. Fritzsche, V.~M.
1173: Shabaev, R.~S. Orts, I.~I. Tupitsyn, Y. Zou, and J. Ullrich, Phys. Rev. Lett.
1174: {\bf 91}, 183001 (2003).
1175:
1176: \end{thebibliography}
1177:
1178: \end{document}
1179:
1180: