1: \documentclass{svmult}
2: % Springer Version [10/24/02 PM]
3: % LA-UR-02-6889
4: %%%%%%%%%%% Author macros begin:
5: %%%%%%%%%%% Author macros end
6: \usepackage{makeidx}
7: \usepackage{graphicx}
8: \usepackage{multicol}
9:
10: \makeindex
11:
12: \begin{document}
13:
14: \title*{The Structure of Light Nuclei and Its Effect on Precise Atomic
15: Measurements}
16: \titlerunning{The Structure of Light Nuclei}
17:
18: \author{James L.\ Friar}
19: \institute{Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM,
20: USA \texttt{friar@lanl.gov}}
21:
22: \label{04}
23: \maketitle
24:
25: \vspace*{-2.5in}
26: \hfill \fbox{\parbox[t]{0.95in}{LA-UR-02-6889}}
27: \vspace*{2.2in}
28:
29: \begin{abstract}
30: This review consists of three parts: (a) what every atomic physicist needs to
31: know about the physics of light nuclei; (b) what nuclear physicists can do for
32: atomic physics; (c) what atomic physicists can do for nuclear physics. A brief
33: qualitative overview of the nuclear force and calculational techniques for light
34: nuclei will be presented, with an emphasis on debunking myths and on recent
35: progress in the field. Nuclear quantities that affect precise atomic
36: measurements will be discussed, together with their current theoretical and
37: experimental status. The final topic will be a discussion of those atomic
38: measurements that would be useful to nuclear physics, and nuclear calculations
39: that would improve our understanding of existing atomic data.
40: \end{abstract}
41:
42: \section{Introduction}
43:
44: \begin{quote}
45: \em ....numerical precision is the very soul of science.....
46: \end{quote}
47:
48: This quote\cite{darcy} from Sir D'Arcy Wentworth Thompson, considered by many to
49: be the first biomathematician, could well serve as the motto of the field of
50: precise atomic measurements, since precision is the {\it raison d'\^etre} of
51: this discipline. I have always been in awe of the number of digits of accuracy
52: achievable by atomic physics in the analysis of simple atomic
53: systems\cite{2S1S}. Nuclear physics, which is my primary field and interest,
54: must usually struggle to achieve three digits of numerical significance, a level
55: that atomic physics would consider a poor initial effort, much less a decent
56: final result.
57:
58: The reason for the differing levels of accuracy is well known: the theory of
59: atoms is QED, which allows one to calculate properties of few-electron systems
60: to many significant figures\cite{review}. On the other hand, no aspect of
61: nuclear physics is known to that precision. For example, a significant part of
62: the ``fundamental'' nuclear force between two nucleons must be determined
63: phenomenologically by utilizing experimental information from nucleon-nucleon
64: scattering\cite{2gen}, very little of which is known to better than 1\%. In
65: contrast to that level of precision, energy-level spacings in few-electron atoms
66: can be measured so precisely that nuclear properties influence significant
67: digits in those energies\cite{d-p}. Thus these experiments can be interpreted
68: as either a measurement of those nuclear properties, or corrections must be
69: applied to eliminate the nuclear effects so that the resulting measurement tests
70: or measures non-nuclear properties. That is the purview of this review.
71:
72: The single most difficult aspect of a calculation for any theorist is assigning
73: uncertainties to the results. This is not always necessary, but in calculating
74: nuclear corrections to atomic properties it is essential to make an effort. That
75: is just another way to answer the question,``What confidence do we have in our
76: results?'' Because it is important for atomic physicists to be able to judge
77: nuclear results to some degree, this discussion has been slanted towards answers
78: to two questions that should be asked by every atomic physicist. The first is:
79: ``What confidence should I have in the values of nuclear quantities that are
80: required to analyze precise atomic experiments?'' The second question is: ``What
81: confidence should I have that the nuclear output of my atomic experiment will be
82: put to good use by nuclear physicists?''
83:
84: \section{Myths of Nuclear Physics}
85:
86: Every field has a collection of myths, most of them being at least partially
87: true at one time. Myths propagate in time and distort the reality of the
88: present. A number of these are collected below, some of which the author once
89: believed. The resolution of these ``beliefs'' also serves as a counterpoint to
90: the very substantial progress made in light-nuclear physics in the past 15
91: years, which continues unabated.
92:
93: My myth collection includes:
94:
95: $\bullet$ The strong interactions (and consequently the nuclear force) aren't
96: well understood, and nuclear calculations are therefore unreliable.
97:
98: $\bullet$ Large strong-interaction coupling constants mean that perturbation
99: theory doesn't converge, implying that there are no controlled expansions in
100: nuclear physics.
101:
102: $\bullet$ The nuclear force has no fundamental basis, implying that
103: calculations are not trustworthy.
104:
105: $\bullet$ You cannot solve the Schr\"odinger equation accurately because of the
106: complexity of the nuclear force.
107:
108: $\bullet$ Nuclear physics requires a relativistic treatment, rendering a
109: difficult problem nearly intractable.
110:
111: All of these myths had some (even considerable) truth in the past, but today
112: they are significant distortions of our current level of knowledge.
113:
114: \section{The Nuclear Force}
115:
116: Most of the recent progress in understanding the nuclear force is based on a
117: symmetry of QCD\index{QCD}, which is believed to be the underlying theory of the
118: strong interactions (or an excellent approximation to it). It is generally the
119: case that our understanding of any branch of physics is based on a framework of
120: symmetry principles. QCD has ``natural'' degrees of freedom (quarks and gluons)
121: in terms of which the theory has a simple representation. The (strong) chiral
122: symmetry\index{Chiral~symmetry!in~QCD} of QCD results when the quark masses
123: vanish, and is a more complicated analogue of the chiral symmetry that results
124: in QED\index{Chiral~symmetry!in~QED} when the electron mass vanishes. The
125: latter symmetry explains, for example, why (massless or high-energy) electron
126: scattering from a spherical (i.e., spinless) nucleus vanishes in the backward
127: direction.
128:
129: The problem with this attractive picture is that it does not involve the degrees
130: of freedom most relevant to experiments in nuclear physics: nucleons and pions.
131: It is nevertheless possible to ``map'' QCD (expressed in terms of quarks and
132: gluons) into an ``equivalent'' or surrogate theory expressed in terms of nucleon
133: and pion degrees of freedom. This surrogate works effectively only at low
134: energy. The small-quark-mass symmetry limit becomes a small-pion-mass symmetry
135: limit. In general this (slightly) broken-symmetry theory has $m_{\pi} c^2 \ll
136: \Lambda$, where the pion mass is $m_{\pi} c^2 \cong$ 140 MeV and $\Lambda \sim$
137: 1 GeV is the mass scale of QCD bound states (heavy mesons, nucleon resonances,
138: etc.). The seminal work on this surrogate theory, now called chiral perturbation
139: theory\index{Chiral~perturbation~theory} (or $\chi$PT), was performed by Steve
140: Weinberg\cite{QCD}, and many applications to nuclear physics were pioneered by
141: his student, Bira van Kolck\cite{nQCD}. From my perspective they demonstrated
142: two things that made an immediate impact on my understanding of nuclear
143: physics\cite{pc}: (1) There is an alternative to perturbation theory in coupling
144: constants, called ``power counting\index{Power~counting},'' that converges
145: geometrically like $(Q/\Lambda)^N$, where $Q \sim m_{\pi}c^2$ is a relevant
146: nuclear energy scale, and the exponent $N$ is constrained to have $N \geq 0$;
147: (2) nuclear physics mechanisms are severely constrained by the chiral symmetry.
148: These results provide nuclear physics with a well-founded rationale for
149: calculation.
150:
151: \begin{figure} \centering
152: \includegraphics[scale=0.85,bb= 100 347 478 600]{04-01.ps}
153: \caption{Cartoon of the nuclear potential, $V (r)$, showing regions of
154: importance}
155: \end{figure}
156:
157: This scheme divides the nuclear-force regime in a natural way into a long-range
158: part\index{Nuclear~force!long-range} (which implies a low energy, $Q$, for the
159: nucleons) and a short-range\index{Nuclear~force!short-range} part (corresponding
160: to high energy, $Q$, between nucleons). This is indicated in Fig.~(1), which is
161: a cartoon of the potential between two nucleons meant only to indicate
162: significant regions and mechanisms. Since $\chi$PT is effective only at low
163: energies, we expect that only the long-range part of the nuclear force can be
164: treated successfully by utilizing only the pion degrees of freedom. This would
165: be the region with $r > b$. We need to resort to
166: phenomenology\index{Nuclear~phenomenology} (i.e., fitting to nucleon-nucleon
167: scattering data) to treat systematically the short-range part of the interaction
168: ($r < b$).
169:
170: The long-range nuclear force is calculated in much the same way that atomic
171: physics calculates the interactions in an atom using QED. Both are illustrated
172: in Fig.~(2). The dominant interaction between two nucleons is the exchange of a
173: single pion illustrated in Fig.~(2b) (One-Pion-Exchange Potential or
174: ``OPEP''\index{Nuclear~force!OPEP}) and denoted $V_{\pi}$. Its atomic analogue
175: is one-photon exchange in Fig.~(2a) (containing the dominant Coulomb force).
176: Because it is such an important part of the nuclear potential, it is fair to
177: call $V_{\pi}$ the ``Coulomb force'' of nuclear physics. Smaller contributions
178: arise from the two-pion-exchange potential in Fig.~(2e) (called
179: ``TPEP''\index{Nuclear~force!TPEP}), which is the analogue of two-photon
180: exchange between charged particles shown in Fig.~(2d). There is even an analogue
181: of the atomic polarization force in Fig.~(2g), where two electrons
182: simultaneously polarize their nucleus using their electric fields. The nuclear
183: analogue involving three nucleons simultaneously is illustrated in Figs.~(2h)
184: and (2i), and is called a three-nucleon
185: force\cite{3NF}\index{Nuclear~force!three-nucleon}. Although relatively
186: weak compared to $V_{\pi}$ (a few percent), three-nucleon forces play an
187: important role in fine-tuning nuclear energy levels. The final ingredient is an
188: important short-range interaction\index{Nuclear~force!short-range} (which
189: must be determined by phenomenology) shown in Fig.~(2c) that has no direct
190: analogue in the physics of light atoms. Just as one can exchange three photons,
191: three-pion-exchange is possible and is depicted in Fig.~(2f).
192:
193: It is worth recalling that the uncertainty principle tells us that exchanging
194: light particles produces longer-range forces and exchanging heavier particles
195: produces shorter-range forces. Thus OPEP has a longer range than TPEP, as
196: illustrated in Fig.~(1). Many mechanisms have been proposed for the short-range
197: part of the nuclear force, such as heavy-meson exchange, for example. Any
198: meson-exchange mechanism produces singular forces, which are regularized to make
199: them finite. However one chooses to do this, the part of the nuclear force
200: inside $b$ must be adjusted to fit the nucleon-nucleon scattering
201: data\index{Nuclear~phenomenology}, and no individual parameterization of the
202: short-range force is intrinsically superior (i.e., it doesn't matter how you do
203: it).
204:
205: \begin{figure} \centering
206: \includegraphics[scale=0.70]{04-02.eps}
207: \caption{First- and second-order (in $\alpha$, the fine structure constant)
208: atomic interactions resulting from photon exchange are shown in the left-most
209: column, where solid lines are electrons, wiggly lines are photons, and the
210: shaded line is a nucleus. The analogous nuclear interactions resulting from
211: pion exchange are shown in the middle column, where solid lines are nucleons
212: and dashed lines are pions. Nuclear processes involving short-range interactions
213: (shaded vertical areas) are shown in the right-most column, together with a
214: three-pion-exchange interaction}
215: \end{figure}
216:
217: \begin{figure} \centering
218: \includegraphics[totalheight=3.5in]{04-03.epsi}
219: \caption{$^3P_0$ phase shift (in degrees) calculated only with the OPEP tail for
220: $r > b$ (dashed line), and with one (dotted) and three (solid)
221: short-range-interaction terms added. The experimental results are indicated by
222: separate points with error bars\protect\cite{psa}}
223: \end{figure}
224:
225: How all of this works in practice is indicated in Fig.~(3). Imagine that you
226: throw away {\it all} of the nuclear potential inside $r = b$ (with $b$ chosen to
227: be 1.4 fm) in Fig.~(1), keeping only the tail of the force between two nucleons.
228: Now compute a phase shift (the $^3$P$_0$, for example). This very modest physics
229: input predicts the basic shape of the phase shift (dashed line) as a function of
230: energy. This variation with energy is a consequence of the small pion mass
231: (compared to the energy scale in the figure). What is missing in this curve is a
232: smooth (negative) short-range contribution that grows roughly in proportion to
233: the energy. We can fill in the missing short-range interaction inside $r = b$ by
234: adding a potential term specified by one short-range parameter. This produces
235: the dotted line, which is a rather good fit, and adding two more terms (solid
236: line) produces a nearly perfect fit to the experimental results. Fixing the
237: short-range\index{Nuclear~force!short-range} part of the potential looks very
238: much like making an effective-range expansion. All useful physics is specified
239: by a few parameters, and the details are completely unimportant.
240:
241: What are the consequences of exchanging a pion rather than a photon? The
242: pseudoscalar nature of the pion mandates its spin-dependent coupling to a
243: nucleon, and this leads to a dominant tensor force\index{Nuclear~force!tensor}
244: between two nucleons. Except for its radial dependence, the form of $V_{\pi}$
245: mimics the interaction between two magnetic dipoles, as seen in the Breit
246: interaction, for example. Thus we have in nuclear physics a situation that is
247: the converse of the atomic case: a dominant tensor force and a smaller central
248: force. In order to grasp the difficulties that nuclear physicists face, imagine
249: that you are an atomic physicist in a universe where magnetic (not electric)
250: forces are dominant, and where QED can be solved only for long-range forces and
251: you must resort to phenomenology to generate the short-range part of the force
252: between electrons and nuclei.
253:
254: Although this may sound hopeless, it is merely difficult. The key to handling
255: complexities is adequate computing power, and that became routinely available
256: only in the late 1980s or early 1990s. Since then there has been explosive
257: development in our understanding of light nuclei. Underlying all of these
258: developments is an improved understanding of the nuclear force. It is
259: convenient to divide nuclear forces and their history into three distinct time
260: periods.
261:
262: {\it First-generation}\index{Nuclear~force!first-generation} nuclear forces were
263: developed prior to 1993. They all contained the one-pion-exchange force, but
264: everything else was relatively crude. The fits to the nucleon-nucleon scattering
265: data (needed to parameterize the short-range part of that force) were
266: indifferent.
267:
268: {\it Second-generation}\index{Nuclear~force!second-generation} forces were
269: developed beginning in 1993\cite{2gen}. They were more sophisticated and
270: generally very well fit to the scattering data. As an example of how well the
271: fitting worked, the Nijmegen group (which pioneered this sophisticated
272: procedure) allowed the pion mass to vary in the Yukawa function defining
273: $V_{\pi}$\index{Nuclear~force!OPEP}, and then fit that mass. They also allowed
274: different masses for the neutral and charged pions that were being exchanged and
275: found\cite{pi-mass}
276:
277: \begin{equation} m_{\pi^\pm} = 139.4(10)\, {\rm MeV} \, , \end{equation}
278: \begin{equation} m_{\pi^{0}}\, = 135.6(13)\, {\rm MeV} \, , \end{equation} both
279: results agreeing with free pion masses ($m_{\pi^{\pm}} = 139.57018(35)$ MeV and
280: $m_{\pi^0} = 134.9766(6)$ MeV \cite{PDG}). It is both heartening and a bit
281: amazing that the masses of the pions can be determined to better than 1\% using
282: data taken in reactions that have no free pions! This result is the best
283: quantitative proof of the importance of pion degrees of freedom in nuclear
284: physics.
285:
286: {\it Third-generation}\index{Nuclear~force!third-generation} nuclear forces are
287: currently under development. These forces are quite sophisticated and
288: incorporate two-pion exchange\index{Nuclear~force!TPEP}, as well as $V_{\pi}$.
289: All of the pion-exchange forces (including three-nucleon
290: forces)\index{Nuclear~force!three-nucleon} are being generated in accordance
291: with the rules of chiral perturbation theory\index{Chiral~perturbation~theory}.
292: One expects even better fits to the scattering data. This is clearly work in
293: progress, but preliminary calculations and versions have already
294: appeared\cite{3gen}.
295:
296: \section{Calculations of Light Nuclei}
297:
298: \begin{figure}[p]
299: \includegraphics[scale=0.63]{04-04.ps}
300: \caption{GFMC calculations of the binding energies of the levels (labelled by
301: their spins and parities) of light nuclei with as many as ten nucleons. These
302: calculations use a common Hamiltonian and have a numerical uncertainty on
303: the order of 1\%. Heavy shaded lines to the left are calculated energies
304: (with errors), while light shaded lines to the right are experimental
305: energies. The label ``IL2'' refers to the Illinois-2 model of the
306: three-nucleon force that is used in all of the calculations together with
307: the Argonne V$_{18}$ two-nucleon force}
308: \end{figure}
309:
310: Having a nuclear force is not very useful unless one can calculate nuclear
311: properties with it. Such calculations are quite difficult. Until the middle
312: 1980s only the two-nucleon problem had been solved with numerical errors smaller
313: than 1\%. At that time the three-nucleon systems $^3$H and $^3$He were
314: accurately
315: calculated\index{Accurate~nuclear~calculations!trinucleon@$^3$H~and~$^3$He}
316: using a variety of first-generation nuclear-force models\cite{3N}. Soon
317: thereafter the
318: $\alpha$-particle\index{Accurate~nuclear~calculations!alpha@$^4$He} ($^4$He) was
319: calculated by Joe Carlson, who pioneered a technique that has revolutionized our
320: understanding of light nuclei: Green's Function Monte Carlo
321: (GFMC)\cite{GFMC}\index{Green's~Function~Monte~Carlo}.
322:
323: The difficulty in solving the Schr\"odinger equation for nuclei is easily
324: understood, although it was not initially obvious. Nuclei are best described in
325: terms of nucleon degrees of freedom. Nucleons come in two types, protons and
326: neutrons, which have nearly the same masses and can be considered as the up and
327: down components of an ``isospin'' degree of freedom. If one also includes its
328: spin, a single nucleon thus has four internal degrees of freedom. Two nucleons
329: consequently have 16 internal degrees of freedom, which is roughly the number of
330: components in the nucleon-nucleon force (coupling spin, isospin and orbital
331: motion in a very complicated way). To handle this complexity one again requires
332: fast computers, and that is a fairly recent development.
333:
334: The GFMC\index{Green's~Function~Monte~Carlo} technique has been used to solve
335: for all of the bound (and some unbound) states of nuclei with up to 10 nucleons.
336: One member of this collaboration (Steve Pieper\cite{steve}) calculated that the
337: ten-nucleon Schr\"odinger equation requires the solution of more than 200,000
338: coupled second-order partial-differential equations in 27 continuous variables,
339: and this can be accomplished with numerical errors on the order of 1\%! A subset
340: of the results of this impressive calculation are shown in Fig.~(4)\cite{10A}.
341:
342: Although the nucleon-nucleon scattering data alone can predict the binding
343: energy of the deuteron ($^2$H) to within about 1/2\%, the experimental binding
344: energy is used as input data in fitting the nucleon-nucleon potential. The
345: nuclei $^3$H and $^3$He (not shown) are slightly underbound without a
346: three-nucleon force\index{Nuclear~force!three-nucleon}, and that force can be
347: adjusted to remedy the underbinding. This highlights both the dominant nature of
348: the nucleon-nucleon force and the relative smallness of three-nucleon forces,
349: which is nevertheless appropriate in size to account for the small discrepancies
350: that result from using only nucleon-nucleon forces in calculations of nuclei
351: with more than two nucleons.
352:
353: Once the $^3$H binding energy is fixed, the binding energy of $^4$He is then
354: accurately predicted to within about 1\%. The five-nucleon systems (not shown)
355: are unbound, but their properties are rather well reproduced. The six-nucleon
356: systems are also well predicted. There are small problems with more
357: neutron-rich nuclei (compare $^9$Li with $^7$Li or $^8$He with $^6$He or
358: $^4$He), but only 3 adjustable parameters in the three-nucleon force allow
359: several dozen energy levels to be quite well
360: reproduced\cite{10A}\index{Accurate~nuclear~calculations!a10@$A=$ 2-10}. Because
361: nuclei are weakly bound systems, there are large cancellations between the
362: (large) potential and (large) kinetic energies, leaving small binding energies.
363: The results shown in Fig.~(4) are quite remarkable, especially given that small
364: (fractional) discrepancies in the energy components lead to large effects on the
365: binding energies.
366:
367: We note finally that power counting\index{Power~counting} can be used to show
368: that light nuclei are basically non-relativistic, and relativistic corrections
369: are on the order of a few percent. Power counting is a powerful qualitative
370: technique for determining the relative importance of various mechanisms in
371: nuclear physics.
372:
373: \section{What Nuclear Physics Can Do for Atomic Physics}
374:
375: With our recently implemented computational skills we in nuclear physics can
376: calculate many properties of light nuclei with fairly good accuracy. This is
377: especially true for the deuteron, which is almost unbound and is computationally
378: simple. Although nuclear experiments don't have the intrinsic accuracy of
379: atomic experiments, many nuclear quantities that are relevant to precise atomic
380: experiments can also be measured using nuclear techniques, and usually with
381: fairly good accuracy.
382:
383: What quantities are we talking about? The nuclear length
384: scale\index{Nuclear~scales} is set by $R \sim 1$ fm $= 10^{-5}$ \AA. The much
385: larger atomic length scale of $a_0 \sim $ 1 \AA\ means that an expansion in
386: powers of $R/a_0$ makes great sense, and a typical wavelength for an atomic
387: electron is so large compared to the nuclear size that only moments of the
388: nuclear observables come into play. This also corresponds to an expansion in
389: $\alpha$, the fine-structure constant, and $m_{\rm e} R$, where $m_{\rm e}$ is
390: the electron mass. This is a rapidly converging series.
391:
392: For processes that have nuclear states inside loops (such as polarizabilities)
393: the excitation energies of those states play a significant role. Although states
394: of any energy can be excited in principle, in practice the effective energy of
395: (virtual) excitation for light nuclei (call it $\bar{\omega}_N$) is within a
396: factor of two of 10~MeV, except for the more tightly bound $\alpha$-particle,
397: which also has a smaller radius as a consequence. This number follows from
398: the uncertainty principle and the fact that nucleons in a light nucleus
399: have a radius of about 2~fm. The deuteron's weak binding generates the lowest
400: values, which is about 6 MeV for the deuteron's
401: electric\index{Electric~polarizability!deuterium}
402: polarizability\index{Polarizability!electric} ($\alpha_{\rm E} \sim
403: 1/\bar{\omega}_N$). Using the value of $\bar{\omega}_N$ = 10 MeV, we find
404: $\bar{\omega}_N R \sim 1/10$, which is a reasonably small expansion parameter.
405:
406: \begin{table}[ht]
407: \centering
408: \caption{Orders in $\alpha$ where various contributions to the Lamb shift for
409: S-states have been calculated. The label ``f.s.'' denotes a contribution from
410: nuclear finite size or nuclear structure. Once nuclear physics enters a
411: process at a given order, higher orders will also have nuclear corrections. A
412: ``$\mbox{$-$}$'' indicates that although a complete calculation of nuclear
413: contributions has not been made, such contributions are expected. Names refer
414: to the person who first calculated the leading-order term of that type.
415: References and the meanings of other labels are given in the text}
416: \begin{tabular}{l l l l l}
417: \hline \noalign{\smallskip}
418: Process \hspace{0.5in} & $\alpha^2$ & $\alpha^4$ & $\alpha^5$ & $\alpha^6$ \\
419: \noalign{\smallskip}
420: \hline
421: NR Coulomb & Bohr \hspace{0.2in}& f.s. & f.s. & f.s. \\ \hline
422: Rel. Coulomb & & Dirac & & f.s. \\ \hline
423: Recoil & & Darwin \hspace{0.02in}&\mbox{$-$} &\mbox{$-$}\\ \hline
424: Nucl. Structure & & & f.s. &\mbox{$-$}\rule{0in}{2.5ex}\\ \hline
425: Vacuum Pol. & & & Uehling \hspace{0.2in}& f.s. \\ \hline
426: Radiative & & & Bethe & f.s. \\ \hline
427: \end{tabular}
428: \end{table}
429:
430: At what levels do various nuclear mechanisms affect the Lamb
431: shift\index{Lamb~shift!nuclear~finite~size}? The (lowest) orders in $\alpha$
432: that receive nuclear contributions (for S-states) are sketched in Table~(1).
433: The various mechanisms are divided into static Coulomb (both non-relativistic
434: and relativistic), recoil (inverse powers of the nuclear mass, $M$), nuclear
435: structure, vacuum polarization, and radiative processes. The nuclear effects are
436: conveniently divided into two categories: those that directly involve only the
437: properties of the nuclear ground state, and those that involve virtual excited
438: states and are traditionally called ``nuclear structure.'' A radius is a good
439: example of the former, while a polarizability is the prototype of the latter.
440: Calculational techniques are quite different for these two categories.
441:
442: It is beyond the scope of this review to list detailed formulae and extensive
443: references to past work. I strongly recommend the recent review of
444: \cite{review}, which is extremely well organized. An entire section is devoted
445: to nuclear contributions, and these are listed in their Table~(10) with
446: references and numerical values for the hydrogen atom. A sketch of how these
447: contributions scale is given below together with some of the more recent
448: references.
449:
450: The leading-order non-relativistic energy is simply the Bohr energy of order
451: $\alpha^2$. Nuclear finite-size\index{Lamb~shift!nuclear~finite~size}
452: contributions of non-relativistic type (i.e., generated by the Schr\"odinger
453: equation) begin for S-states in order $(Z \alpha)^4$\cite{KKS} and are
454: proportional to $R^2$; they have also been calculated in order $(Z \alpha)^5$
455: and $(Z \alpha)^6$\cite{fs1,fs2}. The Dirac energy has a leading-order $(Z
456: \alpha)^4$ term, while the nuclear finite-size contributions of relativistic
457: type begin in order $(Z \alpha)^6$ and are proportional to $R^2$. A recent
458: calculation exists for deuterium\cite{ho-size}. The non-relativistic finite-size
459: corrections of order $(Z \alpha)^5$ and $(Z \alpha)^6$ are tiny for electronic
460: atoms (they contain higher powers of $m_{\rm e} R$, which is very small), but
461: are not necessarily small for muonic atoms ($m_{\mu} R$ is about 1 for most
462: light nuclei), which was the original motivation for developing them. P-state
463: finite-size effects begin in order $(Z \alpha)^6$ and are of both relativistic
464: ($\sim R^2$) and non-relativistic ($\sim R^4$) types.
465:
466: The most important nuclear-structure mechanism is the electric
467: polarizability\index{Electric~polarizability} (which has a long history and will
468: be discussed in more detail later), and this generates a leading contribution of
469: order $\alpha^2 (Z \alpha)^3$. Coulomb corrections
470: \index{Electric~polarizability!Coulomb~corrections} of order $\alpha^2 (Z
471: \alpha)^4$ were developed in the context of a greatly simplified model of the
472: polarizability in muonic atoms\cite{He4-x} (which would not be applicable to
473: electronic atoms).
474:
475: The Uehling mechanism for vacuum polarization is of order $\alpha (Z \alpha)^4$,
476: while the first nuclear corrections are of order $\alpha (Z
477: \alpha)^5$\cite{fs-vp,fs-vp-h,fs-rad, fs-rad-p}. The leading-order radiative
478: process is also of order $\alpha (Z \alpha)^4$, while the nuclear finite-size
479: corrections begin in order $\alpha (Z \alpha)^5$\cite{fs-rad,fs-rad-p}. Both of
480: these nuclear corrections are proportional to $R^2$. Recoil corrections have a
481: long and interesting history that predates the Schr\"odinger equation (C.\ G.\
482: Darwin derived the leading term of order $(Z \alpha)^4$ using Bohr-Sommerfeld
483: quantization; see the references in \cite{breit}). To the best of my knowledge
484: no published calculation exists for the nuclear-finite-size recoil corrections,
485: which begin in order $(Z \alpha)^5$, although the techniques of \cite{GY} lead
486: to a result proportional to $R^2/M$, which should be very small. We note finally
487: the hadronic vacuum polarization\index{Vacuum~polarization!hadronic}, which
488: (although not nuclear in origin) is generated by the strong
489: interactions\cite{had-vp}.
490:
491: One quantity through which nuclear size manifests itself is the nuclear charge
492: form factor (the Fourier transform of the nuclear ground-state charge
493: density\index{Nuclear~charge~density}, $\varrho$), which is given by
494: \pagebreak
495:
496: \begin{equation}
497: F (\vec{q}) = \int d^3 r \, \varrho (\vec{r}) \, \exp (i \vec{q} \cdot \vec{r})
498: \cong Z (1- \frac{\vec{q}^2}{6}\langle r^2 \rangle_{\rm ch} + \cdots \, )
499: - \textstyle{1\over2} \, \vec{q}^{\alpha} \vec{q}^{\beta}\, Q^{\alpha \beta} +
500: \cdots \, ,
501: \end{equation}
502: where $\vec{q}$ is the momentum transferred from an electron to the nucleus,
503: $Q^{\alpha \beta}$ is the nuclear quadrupole-moment tensor, $Z$ is the total
504: nuclear charge, and $\langle r^2 \rangle_{\rm ch}$ is the mean-square radius of
505: the nuclear charge density. These moments should dominate the nuclear
506: corrections to atomic energy levels because $|\vec{q}|$ in an atom is set by
507: the (very small) atomic scales. Using $F$ to construct the electron-nucleus
508: Coulomb interaction, one obtains
509:
510: \begin{equation}
511: V_{\rm C} ( \vec{r} ) \cong - \frac{Z \alpha}{r} + \frac{2 \pi Z \alpha}{3}
512: \langle r^2 \rangle_{\rm ch} \, \delta^3 (\vec{r}) - \frac{Q \alpha}{2 r^3}
513: \frac{( 3 \, (\vec{S} \cdot \hat{\vec{r}})^2 -\vec{S}^2)}{S (2 S-1)} +
514: \cdots \, ,
515: \end{equation}
516: where $\vec{S}$ is the nuclear spin operator and $Q$ is the nuclear quadrupole
517: moment\index{Nuclear~quadrupole~moment!Coulomb~interaction} (which vanishes
518: unless the nucleus has spin $S \geq 1$). The Fourier transform of the nuclear
519: ground-state current density\index{Nuclear~current~density} has a similar
520: expansion
521:
522: \begin{equation}
523: \vec{J} ( \vec{q} ) = \int d^3 r \ \vec{J} (\vec{r}) \, \exp ( i \vec{q}
524: \cdot \vec{r} ) \cong -i \vec{q} \times \mbox{\boldmath $\mu$} \, (1-
525: \frac{\vec{q}^2}{6} \langle r^2 \rangle_{\rm M}\, + \, \cdots \, ) \; + \,
526: \cdots \, ,
527: \end{equation}
528: where $\mbox{\boldmath{$\mu$}}$ is the nuclear magnetic
529: moment\index{Magnetic~moment!nuclear} and $\langle r^2 \rangle_{\rm M}$ is the
530: mean-square radius\index{Nuclear~radii} of the magnetization
531: density\index{Nuclear~radii!magnetic}. The first term generates the usual atomic
532: hyperfine interaction.
533:
534: \begin{table}
535: \centering
536: \caption{Values of the root-mean-square charge and magnetic radii and the
537: quadrupole moment (if nonvanishing) of the nucleons and various light nuclei
538: obtained by nuclear experiments, together with a selected reference. If two
539: values are given, the second value is that obtained by an atomic or molecular
540: measurement}
541: \begin{tabular}{l l l l l l l}
542: \hline \noalign{\smallskip}
543: Nucleus \hspace{0.1in}
544: &$\langle r^2 \rangle_{\rm ch}^{1/2}\, ({\rm fm})$\hspace{0.1in}
545: &ref.\hspace{0.3in}
546: & $\langle r^2 \rangle_{\rm M}^{1/2}\, ({\rm fm})$ \hspace{0.2in}
547: &ref. \hspace{0.15in}
548: &Q\, (fm$^2$) \hspace{0.3in}
549: & ref. \\ \noalign{\smallskip} \hline
550: { }H & 0.880\,(15)&\cite{Coulomb}&0.836\,(9)&\cite{fit} &\mbox{$-$} &
551: \rule{0in}{2.5ex} \\
552: & 0.883\,(14)&\cite{2loop} & & &\mbox{$-$} & \\
553: \hline
554: $^2$H & 2.130\,(10)&\cite{ST} & 2.072\,(18)&\cite{ingo}&0.282\,(19) &
555: \cite{Qnuc} \rule{0in}{2.5ex} \\
556: & & & & &0.2860\,(15)&
557: \cite{Qhd1,Qhd2} \rule{0in}{2.5ex} \\ \hline
558: $^3$H & 1.755\,(87)&\cite{ingo} & 1.84\,(18)&\cite{ingo} &\mbox{$-$} &
559: \rule{0in}{2.5ex} \\ \hline
560: $^3$He& 1.959\,(34)&\cite{ingo} & 1.97\,(15)&\cite{ingo} &\mbox{$-$} &
561: \rule{0in}{2.5ex} \\
562: & 1.954\,(8) &\cite{shiner}& & &\mbox{$-$} & \\
563: \hline
564: $^4$He& 1.676\,(8) &\cite{sickx} & \mbox{$-$}& &\mbox{$-$} &
565: \rule{0in}{2.5ex}\\ \hline \hline
566: \noalign{\smallskip}
567: Nucleon & $\langle r^2 \rangle_{\rm ch}^{ }\, ({\rm fm}^2)$ &ref. &
568: $\langle r^2 \rangle_{\rm M}^{1/2}\, ({\rm fm})$ &ref. & \rule{0in}{2.5ex} \\
569: \noalign{\smallskip} \hline
570: n &\mbox{$-$}0.1140\,(26)&\cite{neutron}&0.873\,(11)&\cite{nmag}& &
571: \\ \hline
572: \end{tabular}
573: \end{table}
574:
575: Electron-nucleus scattering experiments\index{Electron-nucleus~scattering} are
576: the primary technique used to measure moments of nuclear charge and current
577: densities that are relevant to atomic physics\cite{ingo}, and some appropriate
578: values of these quantities are tabulated in Table~(2). An exception is the
579: measurement of the deuteron's quadrupole
580: moment\index{Nuclear~quadrupole~moment!deuterium} ($Q = 0.282(19)$ fm$^2$)
581: obtained by scattering polarized deuterons from a high-Z nuclear target
582: at low energy\cite{Qnuc}. This result is consistent with the molecular
583: determination ($Q = 0.2860(15)$ fm$^2$)\cite{Qhd1,Qhd2}, but its error is an
584: order of magnitude larger. Although there is no reason to believe that the
585: (tensor) electric polarizability\index{Electric~polarizability!tensor} of the
586: deuteron\cite{tau} plays a significant role in the H-D (molecular)
587: quadrupole-hyperfine splitting\index{Hyperfine~splitting!quadrupole} that was
588: used to determine $Q$, that correction was not included in the analysis. It was
589: included in the analysis of the nuclear measurement.
590:
591: I highly recommend the recent review of electron-nucleus
592: scattering\index{Electron-nucleus~scattering} by Ingo Sick\cite{ingo}, which
593: contains values of the charge and magnetic radii of light nuclei. That review
594: not only lists the best and most recent values of quantities of interest, but
595: discusses reliability and technical details for those who are interested. One
596: result from that review is listed in Table~(2) and is important for the
597: discussion below. The errors of the tritium ($^3$H)
598: radii\index{Nuclear~radii!3H@$^3$H} are nearly an order of magnitude larger than
599: those of deuterium. Of all the light nuclei tritium is the most poorly known
600: experimentally, although the charge radius can now be calculated with reasonable
601: accuracy.
602:
603: \begin{figure}
604: \centering
605: \includegraphics[scale=0.9]{04-05.ps}
606: \caption{The direct two-photon process is shown in (a), the crossed-photon
607: process in (b), and ``seagull'' contributions in (c). The seagulls reflect
608: non-nucleonic processes and terms necessary for gauge invariance. In these
609: graphs the double lines represent a nucleus, the single lines an electron,
610: the wiggly lines a (virtual) photon. The shading represents the set of
611: nuclear excited states. The loop momentum is $q$, and integrating over this
612: momentum sets the scales of the nuclear part of the process}
613: \end{figure}
614:
615: In addition to moments of the nuclear charge and current densities, various
616: components and moments of the nuclear Compton
617: amplitude\index{Compton~amplitude!nuclear} can play a significant role.
618: Mechanisms that contribute to the polarizabilities are shown in Fig.~(5). The
619: direct (sequential) exchange of photons and the crossed-photon process are shown
620: in (a) and (b), while the ``seagull'' process is shown in (c). The latter
621: mechanism is required by gauge invariance in any model of hadrons with
622: structure. The exchange of pions between nucleons generates such terms, for
623: example\cite{c-mec}\index{Compton~amplitude!meson-exchange~currents}. Because
624: these are loop diagrams, they involve an integral over all momenta ($q$), and
625: this sets the nuclear scales of the problem. The nuclear size
626: scale\index{Nuclear~scales} ($R$), the electron mass ($m_{\rm e}$), and the
627: average virtual-excitation energy ($\bar{\omega}_N$, appropriate to the shaded
628: part of the line in (a) and (b) that indicates excited nuclear states) determine
629: the generalized polarizabilities\cite{ho-pol}\index{Polarizability!generalized}.
630: The process is dominated by the usual electric\index{Polarizability!electric}
631: and magnetic\index{Polarizability!magnetic} polarizabilities and their
632: logarithmic
633: modifications\cite{d-pol}\index{Polarizability!logarithmic~modification}.
634:
635: Specific examples are the (scalar) electric polarizability, $\alpha_{\rm E}$,
636: and the nuclear spin-dependent
637: polarizability\index{Polarizability!spin-dependent} ($\sim \vec{S}$). The latter
638: interacts with the electron spin to produce a contribution to the
639: electron-nucleus hyperfine
640: splitting\index{Hyperfine~splitting!spin-dependent~polarizability}. There
641: exists a recent calculation of the latter for
642: deuterium\cite{d-nu}\index{Hyperfine~splitting!deuterium}. Values of the nuclear
643: electric polarizability for light
644: nuclei\index{Electric~polarizability!light~nuclei} obtained from calculations or
645: experiments are listed in Table~(3). There are either calculations or
646: measurements of $\alpha_{\rm E}$ for $^2$H\cite{d-pol,d-pol-g,d-pol-x},
647: $^3$H\cite{3pol} and $^3$He\cite{3pol,4pol,he3-pol,rinker}, and
648: $^4$He\cite{He4-x,He4-t}\index{Lamb~shift!MU@muonic $^4$He}. With the exception
649: of the $^3$He experimental results, there is reasonable consistency. The
650: decreasing size of the polarizabilities for heavier nuclei is caused by their
651: increased binding. The $\alpha$-particle has more than 10 times the binding
652: energy of the deuteron, and its polarizability is an order of magnitude smaller.
653:
654: \begin{table}[ht]
655: \centering
656: \caption{Values of the electric polarizability of light nuclei, both theoretical
657: and experimental, where the latter have been determined by nuclear experiments.
658: No uncertainties were given for the $^3$H, $^3$He, and $^4$He calculations in
659: \cite{3pol,4pol}, but they are likely to be smaller than about 10\%. The
660: $^4$He result was used in analyses of muonic He\protect\cite{He4-x,He4-t}}
661: \begin{tabular}{l l l l l}
662: \hline \noalign{\smallskip}
663: Nucleus \hspace{0.2in}
664: & $\alpha_{\rm E}^{\rm calc}\, ({\rm fm}^3)$ \hspace{0.2in}
665: &ref. \hspace{0.2in}
666: &$\alpha_{\rm E}^{\rm exp}\, ({\rm fm}^3)$ \hspace{0.2in}
667: &ref. \\ \noalign{\smallskip} \hline
668: $^2$H &0.6328\,(17)&\cite{d-pol} & 0.61\,(4) &\cite{d-pol-g} \\
669: & & & 0.70\,(5) &\cite{d-pol-x} \\ \hline
670: $^3$H & 0.139 &\cite{3pol} &\mbox{$-$} & \rule{0in}{2.5ex} \\ \hline
671: $^3$He & 0.145 &\cite{4pol} & 0.250\,(40) &\cite{he3-pol}\\
672: & & & 0.130\,(13) &\cite{rinker} \\
673: \hline
674: $^4$He & 0.076 &\cite{4pol} & 0.072\,(4) &\cite{He4-x}
675: \rule{0in}{2.5ex}\\ \hline
676: \end{tabular}
677: \end{table}
678:
679: The physics of hyperfine splittings is in general rather different from the
680: physics that contributes to the Lamb shift. It should therefore be no surprise
681: that the nuclear physics that contributes to hyperfine splittings is also
682: quite different; it is also more complicated than its Lamb-shift counterpart.
683: The dominant nuclear physics that we discussed previously was the physics of the
684: nuclear charge density\index{Nuclear~charge~density}, in the form of moments of
685: the static charge density (i.e., radii) and electric dipole moments that
686: contribute to polarizabilities.
687:
688: The primary nuclear mechanism in hyperfine splittings is the magnetic
689: interaction caused by the nuclear magnetization
690: density\index{Nuclear~current~density!magnetization density}. This density
691: is not as well understood as the charge density. The primary reason is that
692: the same mesons whose exchange binds nuclei together also contribute to the
693: nuclear currents if they carry a charge. The pions that we discussed earlier
694: generate a very important component of that current\cite{c-mec}. The reason for
695: the dichotomy between nuclear charges and currents can be understood by
696: imagining that charged-meson exchange between nucleons is instantaneous. In this
697: limit we know that the transmitted charge is always on a nucleon. In other words
698: only nucleon degrees of freedom matter, which is the normal situation for the
699: charge density. The power counting\index{Power~counting} that we discussed
700: earlier states that corrections to the nuclear charge operator are small
701: ($\sim$ 1\%), and include a type that vanishes for instantaneous meson
702: exchanges. That is not the case for the current, however, since any flow of
703: charge (even from a virtual meson) produces a current that is not simply related
704: to nucleon degrees of freedom, and that current can couple to photons. These
705: meson-exchange currents\index{Nuclear~current~density!meson-exchange~currents}
706: (often denoted ``MEC'') can be as large as those generated by the usual nuclear
707: convection current. Various tricks can be used to eliminate part of our
708: ignorance, but the nuclear current density is less well understood than the
709: nuclear charge density. Atomic hyperfine splittings provide us with an excellent
710: opportunity to learn about nuclear currents in a very different setting.
711:
712: Although most of the hyperfine experiments in light
713: atoms\index{Hyperfine~splitting!light~atoms} were performed decades ago, there
714: has recently been renewed theoretical interest, and the accuracy of the QED
715: calculations is sufficient to extract nuclear information\cite{sav-nu}. The
716: differences between the QED calculations and the experimental results can be
717: interpreted as nuclear corrections, and those are significant, as indicated in
718: Table~(4). The S-state results in this table (presented as a ratio) have been
719: taken from Table~(1) of the recent work of Ivanov and Karshenboim\cite{sav-nu}.
720:
721: \begin{table}[htb]
722: \centering
723: \caption{Difference between hyperfine experiments and QED hyperfine calculations
724: for the $n\underline{\rm th}$ S-state of light hydrogenic atoms times $n^3$,
725: expressed as parts per million of the Fermi energy. This difference is
726: interpreted as nuclear contributions to the hyperfine
727: splitting\protect\cite{sav-nu}. A negative entry indicates that the theoretical
728: prediction without nuclear corrections is too large}
729: \begin{tabular}{l l l l l}
730: \hline \noalign{\smallskip}
731: \multicolumn{5}{c}{$n^3 (E_{\rm hfs}^{\rm exp} -
732: E_{\rm hfs}^{\rm QED})/E_{\rm F}\, {\rm (ppm)}$} \\ \noalign{\smallskip} \hline
733: State \hspace{0.3in} & H \hspace{0.3in} & $^2$H \hspace{0.3in}
734: & $^3$H \hspace{0.3in} & $^3$He$^+$ \rule{0in}{2.5ex}\\ \hline \hline
735: 1S & -33 & 138 & -38 & 222 \\ \hline
736: 2S & -33 & 137 &\mbox{$-$} & 221 \\ \hline
737: \end{tabular}
738: \end{table}
739:
740: Hyperfine structure is generated by short-range interactions. The dominant Fermi
741: contribution ($E_{\rm F}$) arises from a $\delta$-function, and that produces a
742: dependence on the square of the electron's $n\underline{\rm th}$ S-state wave
743: function at the origin, $|\phi_n (0)|^2$, which is proportional to 1/n$^3$. Most
744: nuclear effects have the same dependence ($\sim 1/n^3$), which has been removed
745: from the results in Table~(4). The 1S and 2S results are seen to be consistent
746: at this level of accuracy, with 1S experimental results typically being much
747: more accurate.
748:
749: More calculations of the nuclear contributions to hyperfine splittings in light
750: atoms\index{Hyperfine~splitting!light~atoms} are badly needed if we are to use
751: this information to learn about the currents in light nuclei. These
752: contributions come in the form of Zemach
753: moments\cite{zemach}\index{Hyperfine~splitting!Zemach~moments} (ground-state
754: quantities) and spin-dependent
755: polarizabilities\index{Hyperfine~splitting!spin-dependent~polarizability}
756: (discussed above). There exists a considerable literature on the latter subject
757: dating back 50 years. The recent work of Mil'shtein and Khriplovich\cite{d-nu}
758: has pointed out a serious defect in that older work. Although the leading-order
759: terms are essentially non-relativistic in origin (for the nucleons in a
760: nucleus), the sub-leading-order terms are not, and require relativity (for the
761: nucleons) in order to obtain a correct result. This is not terribly surprising,
762: since the same physics that enters that polarizability also enters the
763: Gerasimov-Drell-Hearn sum rule\cite{GDH}\index{Gerasimov-Drell-Hearn~sum~rule},
764: which requires relativity at the nucleon level\cite{GDH-nuc}, and is a topic of
765: considerable current interest in nuclear physics\cite{henry}. The calculations
766: of \cite{d-nu} suggest that deuterium\index{Hyperfine~splitting!deuterium} at
767: least can be understood using fairly simple nuclear models. This needs to be
768: checked using more sophisticated models. We note that the Zemach
769: correction\cite{zemach}\index{Hyperfine~splitting!Zemach~moments} adds to the
770: ratio in Table~(4), improving the agreement between experiment and theory for H
771: and $^3$H. The large positive value of that ratio for deuterium suggests a large
772: polarizability correction, which is confirmed by \cite{d-nu}.
773:
774: \section{The Proton Size}
775:
776: One recurring problem in the hydrogen Lamb shift is the appropriate value of the
777: mean-square radius\index{Nuclear~radii!proton} of the proton, $\langle r^2
778: \rangle_{\rm p}$, to use in calculations. Some older determinations\cite{HMW}
779: disagree strongly with more recent ones\cite{Simon}. As shown in (3), the slope
780: of the charge form factor (with respect to $\vec{q}^2$) at $\vec{q}^2$ = 0
781: determines that quantity. The form factor is measured by scattering electrons
782: from the proton at various energies and scattering
783: angles\index{Electron-nucleus~scattering}.
784:
785: There are (at least) four problems associated with analyzing the
786: charge-form-factor data to obtain the proton size. The first is that the
787: counting rates in such an experiment are proportional to the flux of electrons
788: times the number of protons in the target seen by each electron. That product
789: must be measured. In other words the measured form factor for low $\vec{q}^2$
790: is ($a - b \frac{\vec{q}^2}{6} + \cdots$), where $b/a = \langle r^2 \rangle_{\rm
791: p}$. The measured normalization $a$ (not exactly equal to 1) clearly influences
792: the value and error of $\langle r^2 \rangle_{\rm p}$. Most analyses
793: unfortunately don't take the normalization fully into account, and \cite{norm}
794: estimates that a proper treatment of the normalization of available data could
795: increase $\langle r^2 \rangle_{\rm p}^{1/2}$ by about 0.015 fm and increase its
796: error, as well. In an atom, of course, the normalization is precisely
797: computable.
798:
799: Another source of error is neglecting higher-order corrections in $\alpha$
800: (i.e., Coulomb
801: corrections)\index{Electron-nucleus~scattering!Coulomb~corrections}. and
802: \cite{Coulomb} demonstrates that this increases $\langle r^2 \rangle_{\rm
803: p}^{1/2}$ by about 0.010 fm. A similar problem in analyzing deuterium data was
804: resolved in \cite{ST}. Another difficulty that existed in the past was a lack of
805: high-quality low-$\vec{q}^2$ data. The final problem is that one must use a
806: sufficiently flexible fitting function to represent $F ( \vec{q})$, or the
807: errors in the radius will be unrealistically low. All of the older analyses had
808: one or more of these flaws.
809:
810: Most of the recent analyses\cite{Simon,Coulomb,fit} are compatible if the
811: appropriate corrections are made. An analysis by Rosenfelder\cite{Coulomb}
812: contains all of the appropriate ingredients, and he obtains $\langle r^2
813: \rangle_{\rm p}^{1/2}$ = 0.880(15) fm. There is a PSI experiment now underway to
814: measure the Lamb shift in muonic hydrogen\index{Lamb~shift!muonic~hydrogen},
815: which would produce the definitive result for $\langle r^2 \rangle_{\rm p}$
816: \cite{PSI,savely}. One expects the results of that experiment to be compatible
817: with Rosenfelder's result. Extraction of the proton radius\cite{2loop} from the
818: electronic Lamb shift\index{Lamb~shift!nuclear~finite~size} is now somewhat
819: uncertain because of controversy involving the two-loop diagrams. These diagrams
820: are significantly less important in muonic hydrogen, where the relative roles of
821: the vacuum polarization and radiative processes are reversed.
822:
823: \section{What Atomic Physics Can Do for Nuclear Physics}
824:
825: The single most valuable gift by atomic physics to the nuclear physics community
826: would be the accurate determination of the proton mean-square radius: $\langle
827: r^2 \rangle_{\rm p}$. This quantity is important to nuclear theorists who wish
828: to compare their nuclear wave function calculations with measured mean-square
829: radii. In order for an external source of electric field (such as a passing
830: electron) to probe a nucleus, it is first necessary to ``grab'' the proton's
831: intrinsic charge distribution\index{Nuclear~radii!proton}, which then maps out
832: the mean-square radius of the proton probability
833: distribution\index{Nuclear~radii!wave~function} in the wave function: $\langle
834: r^2 \rangle_{\rm wfn}$. Thus the measured mean-square radius of a nucleus,
835: $\langle r^2 \rangle$, has the following components:
836:
837: \begin{equation}
838: \langle r^2 \rangle = \langle r^2 \rangle_{\rm wfn} + \langle r^2 \rangle_{\rm
839: p}+ \frac{N}{Z} \langle r^2 \rangle_{\rm n} + \frac{1}{Z} \langle r^2
840: \rangle_{\ldots} \, ,
841: \end{equation}
842: where the intrinsic contribution of the N neutrons has been included as well as
843: that of the Z protons, and $\langle r^2 \rangle_{\ldots}$ is the contribution of
844: everything else, including the very interesting (to nuclear physicists)
845: contributions from strong-interaction mechanisms and relativity in the nuclear
846: charge density\cite{czech}\index{Nuclear~charge~density!exotic~contributions}.
847: Because the neutron looks very much like a positively charged core surrounded by
848: a negatively charged cloud, its mean-square radius\index{Nuclear~radii!neutron}
849: has the opposite sign to that of the proton, whose core is surrounded by a
850: positively charged cloud. It should be clear from (6) that $\langle r^2
851: \rangle_{\rm p}$ (which is much larger than $\langle r^2 \rangle_{\rm n}$) is an
852: important part of the overall mean-square radius. Its present uncertainty
853: degrades our ability to test the wave functions of light nuclei.
854:
855: \begin{figure}
856: \centering
857: \includegraphics[scale=1.0]{04-06.eps}
858: \caption{Cartoon of the $^2$H - H isotope shift, illustrating how the effect
859: of the finite size of the proton (shaded small circle) in deuterium is
860: cancelled in the measurement. The finite size of the neutron (open small
861: circle) and the electromagnetic interaction mediated by the strong-interaction
862: (binding) mechanism (indicated by the jagged line between the nucleons) do
863: affect the deuteron's charge radius (see text)}
864: \end{figure}
865:
866: The next most important measurements are isotope
867: shifts\index{Isotope~shift!light~atoms} in light atoms or ions. Since isotope
868: shifts measure differences in frequencies for fixed nuclear charge Z, the effect
869: of the protons' intrinsic size cancels in the difference. This is particularly
870: important given the current lack of a precise value for the proton's radius. The
871: neutrons' effect is relatively small and can be rather easily eliminated, and
872: thus one is directly comparing differences in wave functions, or of small
873: contributions from $\langle r^2 \rangle_{\ldots}$. Isotope shifts are therefore
874: especially ``theorist-friendly'' measurements, since they are closest to
875: measuring what nuclear theorists actually calculate.
876:
877: Precise isotope-shift measurements have been performed for $^4$He -
878: $^3$He\cite{shiner}\index{Isotope~shift!helium~isotopes} and for $^2$H - $^1$H
879: (D-H)\cite{d-p}\index{Isotope~shift!hydrogen~isotopes}. A measurement of $^6$He
880: - $^4$He is being undertaken\cite{ANL} at ANL. Gordon Drake has written about
881: and strongly advocated such measurements in the Li
882: isotopes\cite{Li-IS}\index{Isotope~shift!lithium~isotopes}. These are all highly
883: desirable measurements. Because the $^3$H (tritium) charge radius currently has
884: large errors, in my opinion the single most valuable measurement to
885: be undertaken for nuclear physics purposes would be the tritium-hydrogen ($^3$H
886: - $^1$H) isotope shift. An extensive series of calculations using
887: first-generation nuclear forces found $\langle r^2 \rangle_{\rm wfn}^{1/2}$ for
888: tritium to be 1.582(8) fm\index{Nuclear~radii!3H@$^3$H}, where the
889: ``error'' is a subjective estimate\cite{radius}. This number could likely be
890: improved by using second-generation nuclear forces, although it will never be as
891: accurate as the corresponding deuteron value, which we discuss next.
892:
893: The D-H isotope shift\index{Isotope~shift!deuterium-hydrogen} in the 2S-1S
894: transition reported by the Garching group\cite{d-p} was
895:
896: \begin{equation}
897: \varDelta \nu = 670 \ \, 994 \ \, 334.64 (15) \ {\rm kHz} \, .
898: \end{equation}
899: \hspace*{1.71in} $\uparrow$ \hspace{0.25in} $\uparrow$ \hspace{0.05in}
900: $\uparrow$ \hspace{0.01in} $\uparrow$ \\ Most of this effect is due to the
901: different masses of the two isotopes (and begins in the first significant
902: figure, indicated by an arrow). The precision is nevertheless sufficiently high
903: that the mean-square-radius\index{Nuclear~radii!deuterium} effect in the sixth
904: significant figure (second arrow) is much larger than the error. The electric
905: polarizability of the deuteron\index{Electric~polarizability!deuterium}
906: influences the eighth significant figure, while the deuteron's magnetic
907: susceptibility\index{Magnetic~susceptibility!deuterium} contributes to the tenth
908: significant figure. It becomes difficult to trust the interpretation of the
909: nuclear physics at about the 1 kHz level, so improving this measurement probably
910: wouldn't lead to an improved understanding of the nuclear physics.
911:
912: Analyzing this isotope shift and interpreting the residue (after applying all
913: QED corrections) in terms of the deuteron's radius leads to the
914: results\cite{iso} in Table~(5). The very small binding energy of the deuteron
915: produces a long wave function tail outside the nuclear potential (interpretable
916: as a proton cloud around the nuclear center of mass), which in turn leads to an
917: easy and very accurate calculation of the mean-square radius of the (square of
918: the) wave function\index{Nuclear~radii!wave~function}. Subtracting this
919: theoretical radius from the experimental deuteron radius (corrected for the
920: neutron's size\index{Nuclear~radii!neutron}) determines the effect of $\langle
921: r^2 \rangle_{\ldots}$ on the radius. Although this difference is quite small, it
922: is nevertheless significant and half the size of the error in the corresponding
923: electron-scattering measurement (see Table~(2)). The high-precision analysis in
924: Table~(5) of the content of the deuteron's charge radius would have been
925: impossible without the precision of the atomic D-H isotope-shift measurement.
926: This measurement has given nuclear physics unique insight into small mechanisms
927: that are at present poorly understood\cite{MEC}.
928:
929: \begin{table}[htb]
930: \centering
931: \caption{Theoretical and experimental deuteron radii for pointlike nucleons.
932: The deuteron wave function radius corresponding to second-generation nuclear
933: potentials and the experimental point-nucleon charge radius of the deuteron
934: (i.e., with the neutron charge radius removed) are shown in the first two
935: columns, followed by the difference of experimental and theoretical results. The
936: difference of the experimental radius with and without the neutron's size is
937: given last for comparison purposes\protect\cite{neutron}}
938: \begin{tabular}{l l l l}
939: \hline \noalign{\smallskip}
940: $ \langle r^2 \rangle_{\rm wfn}^{1/2}\, ({\rm fm})$ \hspace{0.1in}
941: & \rule{0in}{2.5ex} $_{\rm exp}\langle r^2 \rangle_{\rm pt}^{1/2}\, ({\rm fm})$
942: \hspace{0.2in}
943: &${\rm difference}\, ({\rm fm})$ \hspace{0.2in}
944: &$\varDelta \langle r^2 \rangle_{\rm n}^{1/2}\, ({\rm fm})$ \\
945: \noalign{\smallskip}
946: \hline
947: 1.9687(18) \rule{0in}{2.5ex}& 1.9753(10) & 0.0066(21) & \mbox{$-$}0.0291(7)\\
948: \hline
949: \end{tabular}
950: \end{table}
951:
952: \section{Summary and Conclusions}
953:
954: Nuclear forces and nuclear calculations in light nuclei are under control in a
955: way never before attained. This progress has been possible because of the great
956: increase in computing power in recent years. Many of the nuclear quantities that
957: contribute to atomic measurements have been calculated or measured to a
958: reasonable level of accuracy, a level that is improving with time. Isotope
959: shifts are valuable contributions to nuclear physics knowledge, and are
960: especially useful to theorists who are interested in testing the quality of
961: their wave functions for light nuclei. In special cases such as deuterium these
962: measurements provide the only insight into the size of small contributions to
963: the electromagnetic interaction that are generated by the underlying
964: strong-interaction mechanisms. In my opinion the tritium-hydrogen isotope
965: shift\index{Isotope~shift!tritium-hydrogen} would be the most useful measurement
966: of that type. One especially hopes that the ongoing PSI experiment is successful
967: in measuring the proton size via the Lamb shift in muonic hydrogen.
968:
969: \section{Acknowledgements}
970: The work of J.\ L.\ Friar was performed under the auspices of the United States
971: Department of Energy. The author would like to thank Savely Karshenboim for his
972: interest in nuclear aspects of precise atomic measurements and for giving me the
973: opportunity to discuss this problem.
974:
975: \begin{thebibliography}{999}
976:
977: \bibitem{darcy} D'Arcy Wentworth Thompson: {\it On Growth and Form} (Cambridge
978: Univ. Press, Cambridge 1959), p. 2
979:
980: \bibitem{2S1S} J.\ Reichert, M.\ Niering, R.\ Holzwarth, M.\ Wietz, Th.\ Udem,
981: and T. W. H\"{a}nsch: Phys.\ Rev.\ Lett.\ {\bf 84}, 3232 (2000) The H 2S-1S
982: frequency was measured to 14 significant figures in this work
983:
984: \bibitem{review} M.\ I.\ Eides, H.\ Grotch, and V.\ A.\ Shelyuto:
985: Phys.\ Rep.\ {\bf 342}, 63 (2001) This recent, comprehensive, and very
986: well-organized review is highly recommended
987:
988: \bibitem{2gen} V.\ G.\ J.\ Stoks, R.\ A.\ M.\ Klomp,
989: C.\ P.\ F.\ Terheggen, and J.\ J.\ de Swart: Phys.\ Rev.\ C {\bf 49}, 2950
990: (1994); R.\ B.\ Wiringa, V.\ G.\ J.\ Stoks, and R.\ Schiavilla: Phys.\
991: Rev.\ C {\bf 51}, 38 (1995); J.\ L.\ Friar, G.\ L.\ Payne, V.\ G.\ J.\ Stoks,
992: and J.\ J.\ de~ Swart: Phys.\ Lett.\ B{\bf 311}, 4 (1993)
993:
994: \bibitem{d-p} A.\ Huber, Th.\ Udem, B.\ Gross, J.\ Reichert, M.\ Kourogi,
995: K.\ Pachucki, M.\ Weitz, and T.\ W.\ H\"{a}nsch: Phys.\ Rev.\ Lett.\ {\bf 80},
996: 468 (1998)
997:
998: \bibitem{QCD} S.\ Weinberg: Physica A{\bf 96}, 327 (1979); in {\it A
999: Festschrift for I.\ I.\ Rabi}, Transactions of the N.\ Y.\ Academy of Sciences
1000: {\bf 38}, 185 (1977); Nucl.\ Phys.\ B{\bf 363}, 3 (1991); Phys.\ Lett.\
1001: B{\bf 251}, 288 (1990); Phys.\ Lett.\ B{\bf 295}, 114 (1992)
1002:
1003: \bibitem{nQCD} U.\ van Kolck: Thesis, University of Texas, (1993);
1004: C.\ Ord\'o\~nez and U.\ van~Kolck: Phys.\ Lett.\ B{\bf 291}, 459 (1992);
1005: C.\ Ord\'o\~nez, L.\ Ray, and U.\ van Kolck: Phys.\ Rev.\ Lett.\ {\bf 72},
1006: 1982 (1994); U.\ van Kolck: Phys. Rev. C {\bf 49}, 2932 (1994); C.\ Ord\'o\~nez,
1007: L.\ Ray, and U.\ van Kolck: Phys.\ Rev.\ C {\bf 53}, 2086 (1996)
1008:
1009: \bibitem{pc} J.\ L.\ Friar: Few-Body Systems {\bf 22}, 161 (1997)
1010:
1011: \bibitem{3NF} S.\ A.\ Coon, M.\ D.\ Scadron, P.\ C.\ McNamee, B.\ R.\ Barrett,
1012: D.\ W.\ E.\ Blatt, and B.\ H.\ J.\ McKellar: Nucl.\ Phys.\ A{\bf 317},
1013: 242 (1979)
1014:
1015: \bibitem{psa} V.\ G.\ J.\ Stoks, R.\ A.\ M.\ Klomp, M.\ C.\ M.\ Rentmeester,
1016: and J.\ J.\ de Swart: Phys. Rev. C {\bf 48}, 792 (1993)
1017:
1018: \bibitem{pi-mass} R.\ A.\ M.\ Klomp, V.\ G.\ J.\ Stoks, and J.\ J.\ de Swart:
1019: Phys.\ Rev.\ C {\bf 44}, R1258 (1991); V.\ Stoks, R.\ Timmermans,
1020: and J.\ J.\ de Swart: Phys.\ Rev.\ C {\bf 47}, 512 (1993)
1021:
1022: \bibitem{PDG} K.\ Hagiwara, et al.: Phys.\ Rev.\ D {\bf66}, 010001 (2002)
1023:
1024: \bibitem{3gen} M.\ C.\ M.\ Rentmeester, R.\ G.\ E.\ Timmermans, J.\ L.\ Friar,
1025: J.\ J.\ de Swart: Phys.\ Rev.\ Lett. {\bf 82}, 4992 (1999); M.\ Walzl,
1026: U.--G.\ Mei\ss ner, and E.\ Epelbaum: Nucl.\ Phys. A{\bf 693}, 663 (2001);
1027: D.\ R.\ Entem and R.\ Machleidt: Phys.\ Lett.\ B{\bf 524}, 93 (2002)
1028:
1029: \bibitem{3N} C.\ R.\ Chen, G.\ L.\ Payne, J.\ L.\ Friar, and B.\ F.\ Gibson:
1030: Phys.\ Rev.\ C {\bf 31}, 2266 (1985); J.\ L.\ Friar, B.\ F.\ Gibson, and
1031: G.\ L.\ Payne: Phys.\ Rev.\ C {\bf 35}, 1502 (1987)
1032:
1033: \bibitem{GFMC} J.\ Carlson and R.\ Schiavilla: Rev.\ Mod.\ Phys.\ {\bf 70}, 743
1034: (1998) This is a comprehensive review of light nuclei containing an elementary
1035: discussion of GFMC. What Joe Carlson pioneered in the late 1980s was the
1036: application of GFMC to solving the nuclear problem with realistic potentials,
1037: where difficulties of principle exist (the so-called ``fermion problem''). The
1038: GFMC technique was pioneered for nuclear physics with very simplified potentials
1039: in \cite{GFMC-h}. Atomic and molecular applications began about
1040: 1967\cite{GFMC-at}
1041:
1042: \bibitem{GFMC-h} G.\ A.\ Baker, Jr., J.\ L.\ Gammel, B.\ J.\ Hill, and J.\ G.\
1043: Wills: Phys.\ Rev.\ {\bf 125}, 1754 (1962); M.\ H.\ Kalos: Phys.\ Rev.\
1044: {\bf 128}, 1791 (1962)
1045:
1046: \bibitem{GFMC-at} M.\ Kalos: J.\ Comp.\ Phys.\ {\bf 1}, 257 (1967)
1047:
1048: \bibitem{steve} S.\ C.\ Pieper: Private Communication
1049:
1050: \bibitem{10A} S.\ C.\ Pieper, K.\ Varga, and R.\ B.\ Wiringa: Phys.\ Rev.\ C
1051: {\bf 66}, 044310 (2002) This is the latest in a series of calculations of
1052: light nuclei. Details and references for the nuclear forces that they used can
1053: be found here
1054:
1055: \bibitem{KKS} R.\ Karplus, A.\ Klein, and J.\ Schwinger: Phys.\ Rev.\ {\bf 86},
1056: 288 (1952)
1057:
1058: \bibitem{fs1} J.\ L.\ Friar: Phys.\ Lett.\ {\bf 80B}, 157 (1979); J.\ L.\
1059: Friar: Ann.\ Phys.\ (N.Y.) {\bf 122}, 151 (1979)
1060:
1061: \bibitem{fs2} E.\ E.\ Trofimenko: Phys.\ Lett.\ {\bf A73}, 383 (1979);
1062: L.\ A.\ Borisoglebsky and E.\ E.\ Trofimenko: Phys.\ Lett.\ {\bf 81B}, 175
1063: (1979)
1064:
1065: \bibitem{ho-size} J.\ L.\ Friar and G.\ L.\ Payne: Phys.\ Rev.\ A {\bf 56},
1066: 5173 (1997)
1067:
1068: \bibitem{He4-x} J.\ L.\ Friar: Phys.\ Rev.\ C {\bf 16}, 1540 (1977) This work
1069: used the results of \cite{Arkatov}
1070:
1071: \bibitem{Arkatov} Y.\ M.\ Arkatov, et al.: Yad.\ Fiz.\ {\bf 19}, 1172
1072: (1974) [Sov.\ J.\ Nucl.\ Phys.\ {\bf 19}, 598 (1974)]
1073:
1074: \bibitem{fs-vp} J.\ L.\ Friar: Z.\ Phys.\ {\bf A292}, 1 (1979);
1075: (E) {\bf A303}, 84 (1981)
1076:
1077: \bibitem{fs-vp-h} D.\ J.\ Hylton: Phys.\ Rev.\ A {\bf 32}, 1303 (1985)
1078:
1079: \bibitem{fs-rad} M.\ I.\ Eides and H.\ Grotch: Phys.\ Rev.\ A {\bf 56}, R2507
1080: (1997)
1081:
1082: \bibitem{fs-rad-p} K.\ Pachucki: Phys.\ Rev.\ A {\bf 48}, 120 (1993) See the
1083: discussion in \cite{review}
1084:
1085: \bibitem{breit} J.\ L.\ Friar and J.\ W.\ Negele: Phys.\ Lett.\ {\bf 46B}, 5
1086: (1973)
1087:
1088: \bibitem{GY} H.\ Grotch and D.\ R.\ Yennie: Rev.\ Mod.\ Phys.\ {\bf 41}, 350
1089: (1969) This work provides the basis for treating reduced-mass and recoil
1090: corrections
1091:
1092: \bibitem{had-vp} J.\ L.\ Friar, J.\ Martorell, and D.\ W.\ L.\ Sprung:
1093: Phys.\ Rev.\ A {\bf 59}, 4061 (1999) This is only the latest work on this
1094: interesting topic. It contains results of previous calculations together with
1095: references
1096:
1097: \bibitem{Coulomb} R.\ Rosenfelder: Phys. Lett.\ B{\bf 479}, 381 (2000)
1098:
1099: \bibitem{fit} P.\ Mergell, U.--G.\ Mei\ss ner, and D.\ Drechsel:
1100: Nucl.\ Phys.\ A{\bf 596}, 367 (1996)
1101:
1102: \bibitem{2loop} K.\ Melnikov and T.\ van Ritbergen: Phys.\ Rev.\ Lett.\ {\bf 84}
1103: 1673 (2000) They find $\langle r^2 \rangle_{\rm p}^{1/2}$ = 0.883(14) fm
1104:
1105: \bibitem{ST} I.\ Sick and D.\ Trautmann: Phys.\ Lett.\ B{\bf 375}, 16 (1996)
1106:
1107: \bibitem{ingo} I.\ Sick: Prog.\ Part.\ Nucl.\ Phys. {\bf 47}, 245 (2001) This
1108: is an excellent review and is highly recommended
1109:
1110: \bibitem{Qnuc} J.\ E.\ Kammeraad and L.\ D.\ Knutson: Nucl.\ Phys.\
1111: A{\bf 435}, 502 (1985)
1112:
1113: \bibitem{Qhd1} R.\ V.\ Reid, Jr.\ and M.\ L.\ Vaida: Phys.\ Rev.\ Lett.\ {\bf
1114: 29}, 494 (1972); (E) {\bf 34}, 1064 (1975)
1115:
1116: \bibitem{Qhd2} D.\ M.\ Bishop and L.\ M.\ Cheung: Phys. Rev. A {\bf 20},
1117: 381 (1979)
1118:
1119: \bibitem{shiner} D.\ Shiner, R.\ Dixson, and V.\ Vedantham:
1120: Phys.\ Rev.\ Lett.\ {\bf 74}, 3553 (1995) The value deduced in their Ref.~[22]
1121: was used
1122:
1123: \bibitem{sickx} I.\ Sick: Phys.\ Lett.\ B{\bf 116}, 212 (1982)
1124:
1125: \bibitem{neutron} S.\ Kopecky, P.\ Riehs, J.\ A.\ Harvey, and N.\ W.\ Hill:
1126: Phys.\ Rev.\ Lett.\ {\bf 74}, 2427 (1995) The compiled value at the
1127: bottom of Table~1 was used
1128:
1129: \bibitem{nmag} G.\ Kubon, et al.: Phys.\ Lett.\ B{\bf 524}, 26 (2002)
1130:
1131: \bibitem{tau} M.\ H.\ Lopes, J.\ A.\ Tostevan, and R.\ C.\ Johnson:
1132: Phys.\ Rev.\ C {\bf 28}, 1779 (1983)
1133:
1134: \bibitem{c-mec} J.\ L.\ Friar: Phys.\ Rev.\ Lett.\ {\bf 36}, 510 (1976)
1135:
1136: \bibitem{ho-pol} J.\ L.\ Friar and G.\ L.\ Payne: Phys.\ Rev.\ C {\bf 56}
1137: 619 (1997)
1138:
1139: \bibitem{d-pol} J.\ L.\ Friar and G.\ L.\ Payne: Phys.\ Rev.\ C {\bf 55}, 2764
1140: (1997) See the references in this work for earlier calculations of
1141: $\alpha_{\rm E}$
1142:
1143: \bibitem{d-nu} A.\ I.\ Mil'shtein and I.\ B.\ Khriplovich: JETP {\bf 82}, 616
1144: (1996) [Zh.\ Eksp.\ Teor.\ Fiz.\ {\bf 109}, 1146 (1996)]
1145:
1146: \bibitem{d-pol-g} J.\ L.\ Friar, S.\ Fallieros, E.\ L.\ Tomusiak, D.\ Skopik,
1147: and E.\ G.\ Fuller: Phys.\ Rev.\ C {\bf 27}, 1364 (1983)
1148:
1149: \bibitem{d-pol-x} N.\ L.\ Rodning, L.\ D.\ Knutson, W.\ G.\ Lynch, and
1150: M.\ B.\ Tsang: Phys.\ Rev.\ Lett.\ {\bf 49}, 909 (1982)
1151:
1152: \bibitem{3pol} V.\ D.\ Efros, W.\ Leidemann, and G.\ Orlandini: Phys.\ Lett.\
1153: B{\bf 408}, 1 (1997) Two of their calculations of $\alpha_{\rm E}$ for $^3$H
1154: were averaged in our Table~(3)
1155:
1156: \bibitem{4pol} W.\ Leidemann: in {\it Proceedings of the XVIII\underline{th}
1157: European Conference on Few-Body Problems in Physics}, Few-Body Syst.\ (suppl.)
1158: (in press); Private Communication
1159:
1160: \bibitem{he3-pol} F.\ Goeckner, L.\ O.\ Lamm, and L.\ D.\ Knutson:
1161: Phys.\ Rev.\ C {\bf 43}, 66 (1991)
1162:
1163: \bibitem{rinker} G.\ A.\ Rinker: Phys.\ Rev.\ A {\bf 14}, 18 (1976) He estimated
1164: a 10\% error in his $\alpha_{\rm E}$ calculation for $^3$He
1165:
1166: \bibitem{He4-t} R.\ Rosenfelder: Nucl.\ Phys.\ A{\bf 393}, 301 (1983)
1167:
1168: \bibitem{sav-nu} S.\ G.\ Karshenboim and V.\ G.\ Ivanov: Phys.\ Lett.\ B{\bf
1169: 524}, 259 (2002)
1170:
1171: \bibitem{zemach} C.\ Zemach: Phys.\ Rev. {\bf 104}, 1771 (1956)
1172:
1173: \bibitem{GDH} S.\ D.\ Drell and A.\ C.\ Hearn: Phys.\ Rev.\ Lett.\ {\bf 16},
1174: 908 (1966); S.\ B.\ Gerasimov: Sov.\ J.\ Nucl.\ Phys.\ {\bf 2}, 430 (1966)
1175:
1176: \bibitem{GDH-nuc} J.\ L.\ Friar: Phys.\ Rev.\ C {\bf 16}, 1540 (1977)
1177:
1178: \bibitem{henry} H.\ R.\ Weller: in {\it Proceedings of GDH 2000},
1179: ed. by D.\ Drechsel and L. Tiator (World Scientific, Singapore 2001), p. 145
1180:
1181: \bibitem{HMW} L.\ N.\ Hand, D.\ J.\ Miller, and R.\ Wilson: Rev.\ Mod.\ Phys.\
1182: {\bf 35}, 335 (1963) The value of $\langle r^2 \rangle_{\rm p}^{1/2}$ =
1183: 0.805(11) fm is not reliable and should not be used
1184:
1185: \bibitem{Simon} G.\ G.\ Simon, Ch.\ Schmitt, F.\ Borkowski, V.\ H. Walter:
1186: Nucl.\ Phys.\ A{\bf 333}, 381 (1980)
1187:
1188: \bibitem{norm} C.\ W.\ Wong: Int. J. Mod. Phys. E {\bf 3}, 821 (1994)
1189:
1190: \bibitem{PSI} D.\ Taqqu, et al.: Hyperfine Int.\ {\bf 119}, 311 (1999)
1191:
1192: \bibitem{savely} S.\ G.\ Karshenboim: Can.\ J.\ Phys.\ {\bf 77}, 241 (1999)
1193:
1194: \bibitem{czech} J.\ L.\ Friar: Czech. J.\ Phys.\ {\bf 43}, 259 (1993); H.\
1195: Arenh\"ovel: Czech. J.\ Phys.\ {\bf 43}, 207 (1993) These are introductory
1196: articles treating meson-exchange currents (strong-interaction contributions to
1197: electromagnetic currents). Neglecting such currents and other binding effects is
1198: commonly denoted the ``impulse approximation''
1199:
1200: \bibitem{ANL} Z.- T. Lu: Private Communication
1201:
1202: \bibitem{Li-IS} Z.- C.\ Yan and G.\ W.\ F.\ Drake: Phys. Rev. A {\bf 61},
1203: 022504 (2000)
1204:
1205: \bibitem{radius} J.\ L.\ Friar: in {\it XIV${\underline{th}}$ International
1206: Conference on Few-Body Problems in Physics}, ed.\ by F.\ Gross, AIP Conference
1207: Proceedings {\bf 334}, 323 (1995) Theoretical values of $\langle r^2
1208: \rangle_{\rm wfn}^{1/2}$ for $^3$He and $^3$H were obtained from the fits in
1209: Fig.~(1). The $^3$He value of 1.769(5) fm\index{Nuclear~radii!3He@$^3$He} was
1210: used in the talk
1211:
1212: \bibitem{iso} J.\ L.\ Friar, J.\ Martorell, and D.\ W.\ L.\ Sprung: Phys.\ Rev.\
1213: A {\bf 56}, 4579 (1997) A tiny center-of-mass contribution to the deuteron
1214: radius was overlooked in that work, but has been included in Table~(5) above. I
1215: would like to thank Ingo Sick for pointing out the problem
1216:
1217: \bibitem{MEC} A.\ J.\ Buchmann, H.\ Henning, and P.\ U.\ Sauer: Few-Body
1218: Systems {\bf 21}, 149 (1996)
1219:
1220: \end{thebibliography}
1221: \label{04_}
1222: \begin{theindex}
1223:
1224: \item Accurate~nuclear~calculations
1225: \subitem $A=$ 2-10, 9
1226: \subitem $^4$He, 7
1227: \subitem $^3$H~and~$^3$He, 7
1228:
1229: \indexspace
1230:
1231: \item Chiral~perturbation~theory, 3, 7
1232: \item Chiral~symmetry
1233: \subitem in~QCD, 3
1234: \subitem in~QED, 3
1235: \item Compton~amplitude
1236: \subitem meson-exchange~currents, 13
1237: \subitem nuclear, 13
1238:
1239: \indexspace
1240:
1241: \item Electric~polarizability, 11
1242: \subitem Coulomb~corrections, 11
1243: \subitem deuterium, 10, 18
1244: \subitem light~nuclei, 14
1245: \subitem tensor, 13
1246: \item Electron-nucleus~scattering, 13, 16
1247: \subitem Coulomb~corrections, 16
1248:
1249: \indexspace
1250:
1251: \item Gerasimov-Drell-Hearn~sum~rule, 16
1252: \item Green's~Function~Monte~Carlo, 7, 9
1253:
1254: \indexspace
1255:
1256: \item Hyperfine~splitting
1257: \subitem deuterium, 14, 16
1258: \subitem light~atoms, 15, 16
1259: \subitem quadrupole, 13
1260: \subitem spin-dependent~polarizability, 14, 16
1261: \subitem Zemach~moments, 16
1262:
1263: \indexspace
1264:
1265: \item Isotope~shift
1266: \subitem deuterium-hydrogen, 18
1267: \subitem helium~isotopes, 18
1268: \subitem hydrogen~isotopes, 18
1269: \subitem light~atoms, 17
1270: \subitem lithium~isotopes, 18
1271: \subitem tritium-hydrogen, 19
1272:
1273: \indexspace
1274:
1275: \item Lamb~shift
1276: \subitem muonic $^4$He, 14
1277: \subitem muonic~hydrogen, 17
1278: \subitem nuclear~finite~size, 10, 11, 17
1279:
1280: \indexspace
1281:
1282: \item Magnetic~moment
1283: \subitem nuclear, 12
1284: \item Magnetic~susceptibility
1285: \subitem deuterium, 18
1286:
1287: \indexspace
1288:
1289: \item Nuclear~charge~density, 11, 14
1290: \subitem exotic~contributions, 17
1291: \item Nuclear~current~density, 12
1292: \subitem magnetization density, 14
1293: \subitem meson-exchange~currents, 15
1294: \item Nuclear~force
1295: \subitem first-generation, 7
1296: \subitem long-range, 3
1297: \subitem OPEP, 4, 7
1298: \subitem second-generation, 7
1299: \subitem short-range, 3--5
1300: \subitem tensor, 6
1301: \subitem third-generation, 7
1302: \subitem three-nucleon, 4, 7, 9
1303: \subitem TPEP, 4, 7
1304: \item Nuclear~phenomenology, 3, 5
1305: \item Nuclear~quadrupole~moment
1306: \subitem Coulomb~interaction, 12
1307: \subitem deuterium, 13
1308: \item Nuclear~radii, 12
1309: \subitem $^3$H, 13, 18
1310: \subitem $^3$He, 22
1311: \subitem deuterium, 18
1312: \subitem magnetic, 12
1313: \subitem neutron, 17, 19
1314: \subitem proton, 16, 17
1315: \subitem wave~function, 17, 19
1316: \item Nuclear~scales, 10, 14
1317:
1318: \indexspace
1319:
1320: \item Polarizability
1321: \subitem electric, 10, 14
1322: \subitem generalized, 14
1323: \subitem logarithmic~modification, 14
1324: \subitem magnetic, 14
1325: \subitem spin-dependent, 14
1326: \item Power~counting, 3, 9, 15
1327:
1328: \indexspace
1329:
1330: \item QCD, 3
1331:
1332: \indexspace
1333:
1334: \item Vacuum~polarization
1335: \subitem hadronic, 11
1336:
1337: \end{theindex}
1338: \end{document}
1339: