1: \documentclass[aps,pra,twocolumn,showpacs,floatfix]{revtex4}
2: \usepackage{graphicx}% Include figure files
3: \usepackage{dcolumn}% Align table columns on decimal point
4: \usepackage{amsmath}
5: \begin{document}
6:
7: \title{\bf Relativistic effects in two valence electron atoms and ions and
8: search for variation of the fine structure constant}
9: \author{E. J. Angstmann}
10: \author{V. A. Dzuba}
11: \email{V.Dzuba@unsw.edu.au}
12: \author{V. V. Flambaum}
13: \email{V.Flambaum@unsw.edu.au}
14: \affiliation{School of Physics, University of New South Wales,
15: Sydney 2052,Australia}
16:
17: \date{\today}
18:
19: %*****************************************************************
20: \begin{abstract}
21: We perform accurate calculations of the dependence of transition frequencies
22: in two valence electron atoms and ions on a variation of the fine structure
23: constant, $\alpha = e^{2}/\hbar c$. The relativistic Hartree-Fock method is used with many-body
24: perturbation theory and configuration interaction methods to calculate
25: transition frequencies. The results are to be used in atomic-clock-type
26: laboratory experiments designed to test whether $\alpha$ varies in time.
27:
28: \end{abstract}
29: \pacs{PACS: 31.30.Jv, 06.20.Jr 95.30.Dr}
30: \maketitle
31:
32: %*****************************************************************
33: \section{Introduction}
34: Theories unifying gravity with other interactions allow for the possible
35: variation of physical constants (see, e.g. \cite{theo1,theo2,theo3}).
36: Recent analysis of quasar absorption spectra suggests that the fine
37: structure constant $\alpha$ might vary in space-time
38: \cite{quasar1,quasar2,quasar3}. There is an
39: intensive search for alternative ways to test whether $\alpha$
40: is varying. One of the very promising methods to study
41: local present-day variation of fundamental constants in time involves
42: the use of atomic clocks. In particular, optical atomic
43: clock transitions are suitable to study the possible variation of the fine
44: structure constant.
45: This is because the ratio of the frequencies of the optical transitions
46: depend on $\alpha$ alone, while the frequencies of the hyperfine transitions
47: also depend on the nuclear magnetic moments and the electron-proton mass ratio.
48:
49: Laboratory measurements involve measuring how the difference between two
50: frequencies changes with time. To relate a measurement of the change between
51: two frequencies to a change in $\alpha$, the relativistic energy shifts are
52: needed. The relativistic energy shift describes how a level moves as
53: $\alpha$ varies. Two transition frequencies with very different relativistic
54: energy shifts are the most desirable candidates for precision experiments as
55: they will have the largest relative frequency shift between them.
56:
57: The best limit on local present day variation of the fine structure constant
58: published to date was obtained by comparing cesium and rubidium atomic fountain
59: clocks \cite{Marison}. Experiments have also been carried out comparing cesium
60: and magnesium \cite{Godone} and a H-maser compared with a Hg~II
61: clock \cite{Prestage}. There are many proposals for the search of variation of
62: $\alpha$ in atomic optical transitions, some of which were analyzed previously
63: in \cite{Dzuba1,Dzuba2,Dzuba3}. In the present work we perform relativistic
64: many-body calculations to find the relativistic energy shift for many two
65: valence electron atoms and ions. Two valence electron atoms and ions were
66: chosen since many new optical clocks experiments, some of which are currently
67: under construction and some still under consideration, utilize these atoms and
68: ions (e.g.. Al~II \cite{Wineland}, Ca~I \cite{Udem}, Sr~I
69: \cite{Courtillot,Katori,Takamoto}, In~II
70: \cite{Becker,Nagourney,Zanthier}, Yb~I, Hg~I \cite{Bize,Porsev}).
71:
72: \section{Theory}
73:
74: In the present work we perform calculations for closed shell atoms and ions
75: which can also be considered as atoms/ions with two valence electrons above
76: closed shells. We start our calculations from the relativistic Hartree-Fock
77: (RHF) (also known as Dirac-Hartree-Fock) method in the $V^{N}$ approximation.
78: This means that RHF calculations are done for the ground state of the
79: corresponding atom/ion with all electrons included in the self-consistent
80: field. The use of the $V^{N}$ RHF approximation ensures good convergence
81: of the consequent configuration interaction (CI) calculations for
82: the ground state.
83: Good accuracy for excited states is achieved by using a large set of
84: single-electron states. Note that there is an alternative approach
85: which uses the $V^{N-2}$ starting approximation (with two valence electrons
86: removed from the RHF calculations). This approach has some advantages,
87: it is simpler, and ground and excited states are treated equally. However,
88: the convergence with respect to the size of the basis is not as good and the
89: final results are better in the $V^N$ approximation. We use the $V^{N-2}$
90: approximation as a test of the accuracy of calculations of the relativistic
91: energy shifts, while presenting all results in the $V^N$ approximation.
92:
93: We use a form of the single-electron wave function that explicitly includes a
94: dependence on $\alpha$:
95: \begin{equation}
96: \psi(\textbf{r})_{njlm}=\frac{1}{r}\Big(\begin{array}{c}f(r)_{n}
97: \Omega(\textbf{r}/r)_{jlm}\\i\alpha g(r)_{n}\tilde{\Omega}
98: (\textbf{r}/r)_{jlm}\end{array}\Big).
99: \end{equation}
100: This leads to the following form of the RHF equations (in atomic units):
101: \begin{eqnarray}\label{eq:HF}
102: f^{'}_{n}(r)+\frac{\kappa_{n}}{r}f_{n}(r)-[2+\alpha^{2}(\epsilon_{n}-
103: \hat{V}_{HF})]g_{n}(r)=0, \nonumber\\
104: g^{'}_{n}(r)+\frac{\kappa_{n}}{r}g_{n}(r)+(\epsilon_{n}-\hat{V}_{HF})
105: f_{n}(r)=0,
106: \end{eqnarray}
107: where $\kappa=(-1)^{l+j+1/2}(j+1/2)$, $n$ is the principle quantum number
108: and $\hat{V}_{HF}$ is the Hartree-Fock potential. The non-relativistic limit
109: corresponds to setting $\alpha = 0$.
110:
111: We then use the combination of the configuration interaction (CI) method with
112: the many-body perturbation theory (MBPT)\cite{CIMBPT,DJ98}.
113: Interactions between valence electrons are treated using the CI method while
114: correlations between the valence electrons and the core electrons are
115: included by means of the MBPT. We can write the effective CI Hamiltonian
116: for two valence electrons as:
117: \begin{equation}\label{eq:HCI}
118: \hat{H}^{CI} =\hat h_1+\hat h_2+ \hat h_{12}
119: \end{equation}
120: here $\hat{h_{i}}$ ($i=1$ or $2$) is an effective single-electron Hamiltonian given by
121: \begin{equation}\label{eq:hi}
122: \hat{h_{i}}= c \mbox{\boldmath{$\alpha$}}\times {\mathbf{p}} +
123: (\beta-1)mc^{2}-\frac{Ze^{2}}{r_{i}}+\hat V_{core}+\hat \Sigma_{1},
124: \end{equation}
125: $\hat V_{core}$ is the Hartree-Fock potential created by the core electrons,
126: it differs from $\hat V_{HF}$ in Eq. (\ref{eq:HF}) by the contribution of the
127: valence electrons. $\hat \Sigma_{1}$ is the one-electron operator that
128: describes the correlation interaction between a valence electron and the core.
129: The third term in Eq. (\ref{eq:HCI}) describes the interaction of the valence
130: electrons with each other and can be written as
131: \begin{equation}\label{eq:hij}
132: \hat h_{12}=\frac{e^{2}}{r_{12}}+\hat \Sigma_{2}
133: \end{equation}
134: where $\hat{\Sigma_{2}}$ is a two-particle operator that describes the effects of
135: screening of the Coulomb interaction between the valence electrons by the core
136: electrons. The operators $\hat{\Sigma_{1}}$ and $\hat{\Sigma_{2}}$
137: are calculated using the second order of MBPT.
138:
139: We use the same set of single-electron basis states to construct two-electron
140: wave functions for the CI calculations and to calculate $\hat \Sigma$.
141: The set is based on the B-spline technique developed by Johnson {\em et al}
142: \cite{BS1,BS2,BS3}. We use 40 B-splines in a cavity of radius $R=40a_B$ ($a_B$
143: is Bohr radius). The single-electron basis functions are linear combinations
144: of 40 B-splines and are also eigenstates of the Hartree-Fock Hamiltonian
145: (in the $V^N$ potential). Therefore, we have 40 basis functions in each
146: partial wave including the B-spline approximations to the atomic core states.
147: We use a different number of basis states for the CI wave functions and for
148: the calculations of $\hat \Sigma$. Saturation comes much faster for the
149: CI calculations. In these calculations we use 14 states above the core in
150: each partial wave up to $l_{max}=3$. Inclusion of states of higher principal
151: quantum number or angular momentum does not change the result.
152: To calculate $\hat \Sigma$ we use 30 out of 40 states in each partial wave
153: up to $l_{max}=4$.
154:
155: The results for the energies are presented in Table \ref{tab:levels}.
156: We present the energies of the $nsnp$ configuration of two electron atoms/ions
157: with respect to their ground state $^1S_0 \ ns^2$. The states considered
158: for atomic clock experiments are $^3P_0$ and $^3P_1$. However,
159: we present the result for other states as well for completeness,
160: these also make it easier to analyze the accuracy of the calculations.
161: Also, transitions associated with some of these states are observed in quasar
162: absorption spectra (e.g., the $^1S_0 - ^1P_1$ transition in Ca).
163:
164: \begin{table}[bt]
165: \caption{Energies of the $nsnp$ configuration of two electron atoms calculated
166: using $H^{CI}$, $H^{CI}+\hat{\Sigma_{1}}$ and
167: $H^{CI}+\hat{\Sigma_{1}}+\hat{\Sigma_{2}}$; comparison with experiment
168: (cm$^{-1}$)}
169: \begin{ruledtabular}
170: \begin{tabular}{c c c c c c}
171: \label{tab:levels}
172: Atom/ & State & Experiment &\multicolumn{3}{c}{Theory} \\
173: ion & & \cite{Moore} & $\hat H^{CI}$ & $\hat H^{CI}+\hat{\Sigma}_1$ &
174: $\hat H^{CI}+\hat{\Sigma}_{1,2}$ \\
175: \hline
176: AlII & $^{3}P_{0}$ & 37393 & 36403 & 36987 & 37328 \\
177: & $^{3}P_{1}$ & 37454 & 36466 & 37053 & 37393 \\
178: & $^{3}P_{2}$ & 37578 & 36592 & 37185 & 37524 \\
179: & $^{1}P_{1}$ & 59852 & 59794 & 60647 & 60090 \\
180:
181: CaI & $^{3}P_{0}$ & 15158 & 13701 & 14823 & 15011 \\
182: & $^{3}P_{1}$ & 15210 & 13750 & 14881 & 15066 \\
183: & $^{3}P_{2}$ & 15316 & 13851 & 14997 & 15179 \\
184: & $^{1}P_{1}$ & 23652 & 23212 & 24968 & 24378 \\
185:
186: SrI & $^{3}P_{0}$ & 14318 & 12489 & 13897 & 14169 \\
187: & $^{3}P_{1}$ & 14504 & 12661 & 14107 & 14367 \\
188: & $^{3}P_{2}$ & 14899 & 13021 & 14545 & 14786 \\
189: & $^{1}P_{1}$ & 21698 & 20833 & 23012 & 22305 \\
190:
191: InII & $^{3}P_{0}$ & 42276 & 37825 & 39238 & 42304 \\
192: & $^{3}P_{1}$ & 43349 & 38867 & 40394 & 43383 \\
193: & $^{3}P_{2}$ & 45827 & 41168 & 42974 & 45904 \\
194: & $^{1}P_{1}$ & 63034 & 62181 & 64930 & 62325 \\
195:
196: YbI & $^{3}P_{0}$ & 17288 & 14377 & 16352 & 16950 \\
197: & $^{3}P_{1}$ & 17992 & 15039 & 17189 & 17705 \\
198: & $^{3}P_{2}$ & 19710 & 16550 & 19137 & 19553 \\
199: & $^{1}P_{1}$ & 25068 & 24231 & 27413 & 26654 \\
200:
201: HgI & $^{3}P_{0}$ & 37645 & 31864 & 32692 & 37420 \\
202: & $^{3}P_{1}$ & 39412 & 33751 & 34778 & 39299 \\
203: & $^{3}P_{2}$ & 44043 & 38155 & 39781 & 44158 \\
204: & $^{1}P_{1}$ & 54069 & 50247 & 52994 & 56219 \\
205:
206: TlII & $^{3}P_{0}$ & 49451 & 43831 & 43911 & 49865 \\
207: & $^{3}P_{1}$ & 52393 & 47091 & 47350 & 52687 \\
208: & $^{3}P_{2}$ & 61725 & 55988 & 56891 & 62263 \\
209: & $^{1}P_{1}$ & 75660 & 74291 & 76049 & 74717 \\
210: \end{tabular}
211: \end{ruledtabular}
212: \end{table}
213:
214: To demonstrate the importance of the core-valence correlations we
215: include results of pure CI calculations (with no $\hat \Sigma$) as well
216: as the results in which only $\hat \Sigma_1$ is included but $\hat \Sigma_2$
217: is not. One can see that the accuracy of pure CI calculations is about
218: 10\% while inclusion of core-valence correlations improves it significantly
219: to the level of about 1\%. The deviation from experiment of the final
220: theoretical energies for the triplet states of all atoms except Yb is not more
221: than 1\%. For Yb it is 2\%. The accuracy of the singlet states is about 1\%
222: for the ions, 3-4\% for CaI, SrI and HgI and 6\% for YbI. The accuracy of the
223: fine structure intervals ranges from 2 to 7\%. The accuracy of
224: calculations for Yb is not as good as for other atoms because the two electron
225: approximation is a poor approximation for this atom. Electrons from the
226: $4f$ subshell, which are kept frozen in present calculations, are relatively
227: easy to excite and corresponding configurations give substantial
228: contribution to the energy. Note that we do include these excitations
229: perturbatively, into the $\hat \Sigma$ operator. However, due to their
230: large contribution, second-order treatment of the excitations from the $4f$
231: subshell is not very accurate. On the other hand, the CI+MBPT
232: results for Yb are still much better than pure CI values.
233:
234: Note also that the CI+MBPT results presented in Table \ref{tab:levels}
235: are in good agreement with similar calculations in Refs.
236: \cite{Kozlov1,Kozlov2}.
237:
238: \section{Results and Discussion}
239:
240: \begin{table}[tb]
241: \caption{Calculated $q$ coefficients, for transitions from the ground state, using $H^{CI}$, $H^{CI}+\hat{\Sigma_{1}}$
242: and $H^{CI}+\hat{\Sigma_{1}}+\hat{\Sigma_{2}}$}
243: \begin{ruledtabular}
244: \begin{tabular}{c c c c c c}\label{tab:q}
245: Atom/ion & State &
246: \multicolumn{1}{c}{$\hat H^{CI}$} &
247: \multicolumn{1}{c}{$\hat H^{CI}+\hat \Sigma_{1}$} &
248: \multicolumn{1}{c}{$H^{CI}+\hat\Sigma_{1,2}$} & Other \\
249: \hline
250: AlII & $^{3}P_{0}$ & 138 & 142 & 146 & \\
251: & $^{3}P_{1}$ & 200 & 207 & 211 & \\
252: & $^{3}P_{2}$ & 325 & 340 & 343 & \\
253: & $^{1}P_{1}$ & 266 & 276 & 278 & \\
254:
255: CaI & $^{3}P_{0}$ & 108 & 115 & 125 & \\
256: & $^{3}P_{1}$ & 158 & 173 & 180 & 230 \cite{Dzuba1} \\
257: & $^{3}P_{2}$ & 260 & 291 & 294 & \\
258: & $^{1}P_{1}$ & 228 & 238 & 250 & 300 \cite{Dzuba1} \\
259:
260: SrI & $^{3}P_{0}$ & 384 & 396 & 443 & \\
261: & $^{3}P_{1}$ & 560 & 609 & 642 & 667 \cite{Dzuba4} \\
262: & $^{3}P_{2}$ & 939 & 1072 & 1084 & \\
263: & $^{1}P_{1}$ & 834 & 865 & 924 & 1058 \cite{Dzuba4} \\
264:
265: InII & $^{3}P_{0}$ & 3230 & 2932 & 3787 & 4414 \cite{Dzuba3} \\
266: & $^{3}P_{1}$ & 4325 & 4125 & 4860 & 5323 \cite{Dzuba3} \\
267: & $^{3}P_{2}$ & 6976 & 7066 & 7767 & 7801 \cite{Dzuba3} \\
268: & $^{1}P_{1}$ & 6147 & 6103 & 6467 & \\
269:
270: YbI & $^{3}P_{0}$ & 2339 & 2299 & 2714 & \\
271: & $^{3}P_{1}$ & 3076 & 3238 & 3527 & \\
272: & $^{3}P_{2}$ & 4935 & 5707 & 5883 & \\
273: & $^{1}P_{1}$ & 4176 & 4674 & 4951 & \\
274:
275: HgI & $^{3}P_{0}$ & 13231 & 9513 & 15299 & \\
276: & $^{3}P_{1}$ & 15922 & 12167 & 17584 & \\
277: & $^{3}P_{2}$ & 22994 & 19515 & 24908 & \\
278: & $^{1}P_{1}$ & 20536 & 16622 & 22789 & \\
279:
280: TlII & $^{3}P_{0}$ & 14535 & 11101 & 16267 & 19745 \cite{Dzuba3} \\
281: & $^{3}P_{1}$ & 18476 & 14955 & 18845 & 23213 \cite{Dzuba3} \\
282: & $^{3}P_{2}$ & 32287 & 28903 & 33268 & 31645 \cite{Dzuba3} \\
283: & $^{1}P_{1}$ & 28681 & 25160 & 29418 & \\
284:
285: \end{tabular}
286: \end{ruledtabular}
287: \end{table}
288:
289: \begin{table}[tb]
290: \caption{Experimental energies and calculated $q$-coefficients (cm$^{-1}$)
291: for transitions from the ground state $ns^{2}$ to the $nsnp$ configurations of two-electron atoms/ions}
292: \begin{ruledtabular}
293: \begin{tabular}{c c c c l r }\label{tab:final}
294: Atom/Ion & Z & \multicolumn{2}{c}{State} & Energy\cite{Moore} &
295: \multicolumn{1}{c}{q} \\
296: \hline
297: AlII & 13 & $3s3p$ & $^{3}P_{0}$ & 37393.03 & 146 \\
298: & & $3s3p$ & $^{3}P_{1}$ & 37453.91 & 211 \\
299: & & $3s3p$ & $^{3}P_{2}$ & 37577.79 & 343 \\
300: & & $3s3p$ & $^{1}P_{1}$ & 59852.02 & 278 \\
301:
302: CaI & 20 & $4s4p$ & $^{3}P_{0}$ & 15157.90 & 125 \\
303: & & $4s4p$ & $^{3}P_{1}$ & 15210.06 & 180 \\
304: & & $4s4p$ & $^{3}P_{2}$ & 15315.94 & 294 \\
305: & & $4s4p$ & $^{1}P_{1}$ & 23652.30 & 250 \\
306:
307: SrI & 38 & $5s5p$ & $^{3}P_{0}$ & 14317.52 & 443 \\
308: & & $5s5p$ & $^{3}P_{1}$ & 14504.35 & 642 \\
309: & & $5s5p$ & $^{3}P_{2}$ & 14898.56 & 1084 \\
310: & & $5s5p$ & $^{1}P_{1}$ & 21698.48 & 924 \\
311:
312: InII & 49 & $5s5p$ & $^{3}P_{0}$ & 42275 & 3787 \\
313: & & $5s5p$ & $^{3}P_{1}$ & 43349 & 4860 \\
314: & & $5s5p$ & $^{3}P_{2}$ & 45827 & 7767 \\
315: & & $5s5p$ & $^{1}P_{1}$ & 63033.81 & 6467 \\
316:
317: YbI & 70 & $6s6p$ & $^{3}P_{0}$ & 17288.44 & 2714 \\
318: & & $6s6p$ & $^{3}P_{1}$ & 17992.01 & 3527 \\
319: & & $6s6p$ & $^{3}P_{2}$ & 19710.39 & 5883 \\
320: & & $6s6p$ & $^{1}P_{1}$ & 25068.22 & 4951 \\
321:
322: HgI & 80 & $6s6p$ & $^{3}P_{0}$ & 37645.08 & 15299 \\
323: & & $6s6p$ & $^{3}P_{1}$ & 39412.30 & 17584 \\
324: & & $6s6p$ & $^{3}P_{2}$ & 44042.98 & 24908 \\
325: & & $6s6p$ & $^{1}P_{1}$ & 54068.78 & 22789 \\
326:
327: TlII & 81 & $6s6p$ & $^{3}P_{0}$ & 49451 & 16267 \\
328: & & $6s6p$ & $^{3}P_{1}$ & 53393 & 18845 \\
329: & & $6s6p$ & $^{3}P_{2}$ & 61725 & 33268 \\
330: & & $6s6p$ & $^{1}P_{1}$ & 75600 & 29418 \\
331:
332: \end{tabular}
333: \end{ruledtabular}
334: \end{table}
335:
336: In the vicinity of the $\alpha_{0}$, the present day value of $\alpha$, the
337: frequency of a transition, $\omega$, can be written as:
338: \begin{equation}\label{eq:omega}
339: \omega = \omega_{0}+qx,
340: \end{equation}
341: where $x=(\frac{\alpha}{\alpha_{0}})^{2}-1$, $\omega_{0}$ is the present day
342: experimental value of the frequency and the $q$ coefficient is the relativistic
343: energy shift that determines the frequency dependence on $\alpha$. It is clear
344: from the above expression that $q$ coefficients can be described by
345: \begin{equation*}
346: q=\frac{d\omega}{dx}|_{x=0}.
347: \end{equation*}
348: Thus, in order to calculate $q$ coefficients the atomic energy levels of the
349: atoms and ions of interest at different values of $x$ need to be calculated.
350: The relativistic energy shift $q$ is then calculated using the formulae
351: \begin{equation}
352: q=\frac{\omega(\Delta x) - \omega(-\Delta x)}{2\Delta x}
353: \end{equation}
354: and
355: \begin{equation}
356: q=\frac{16(\omega(\Delta x) - \omega(-\Delta x)) -
357: 2(\omega(2\Delta x) - \omega(-2\Delta x))}{24\Delta x}.
358: \end{equation}
359: The second formula is needed to check for non-linear contributions to
360: $d\omega/dx$. We use $\Delta x = 0.1$ and $\Delta x = 0.125$. The results are
361: presented in Table \ref{tab:q}.
362:
363: As for the energies, we use three different approximations to calculate
364: relativistic energy shifts: (1) pure CI approximation for two valence
365: electrons, (2) CI with $\hat \Sigma_1$ and (3) CI+MBPT approximation with both
366: $\hat \Sigma_1$ and $\hat \Sigma_2$ included.
367: Inclusion of core-valence correlations lead to increased values of the
368: $q$-coefficients. This is because the correlation interaction of a valence
369: electron with the core introduces an additional attraction which increase the
370: density of the valence electron in the vicinity of the nucleus and thus
371: emphasize the importance of the relativistic effects.
372:
373: Note that $\hat \Sigma_1$ and $\hat \Sigma_2$ are of the same order
374: and need to be included simultaneously to obtain reliable results.
375: $\hat{\Sigma_1}$ is much easier to calculate and inclusion of $\hat \Sigma_1$
376: alone often leads to significant improvements of the results
377: for the energies (see Table \ref{tab:levels}).
378: However, the results for the $q$-coefficients
379: show that neglecting $\hat \Sigma_2$ may lead to significant loss in
380: accuracy. Indeed, the results for $q$'s with $\hat \Sigma_1$ alone
381: are often smaller than those obtained in pure CI and CI+MBPT approximations
382: and differ from final values by up to 50\%.
383: Since neglecting $\hat \Sigma_2$
384: cannot be justified, we present results without $\hat \Sigma_2$ for
385: illustration purpose only.
386:
387: The accuracy of the calculation of the
388: $q$-coefficients can be estimated by comparing the CI and CI+MBPT results
389: calculated in the $V^{N}$ and $V^{N-2}$ approximations and also by
390: comparing the final results for the energies
391: (including fine structure
392: intervals) with experimental values.
393: As one can see from Table \ref{tab:q} inclusion of the core-valence
394: correlations can change the values of the $q$-coefficients by more than 15\%.
395: However, the accuracy of the energies improves significantly when
396: core-valence correlations are included. It is natural to expect that
397: the final accuracy for the $q$-coefficients is also higher when
398: core-valence correlations are included.
399: Comparison with our previous results also shows some deviation
400: on approximately the same level (the largest relative discrepancy is
401: for Ca where relativistic effects are small and high accuracy is not needed).
402: Most of this discrepancy can be attributed to the inaccuracy of our
403: old, less complete calculations. Comparison
404: between the energies calculated in the $V^{N}$ and $V^{N-2}$ approximations and the experimental values suggest that 10\% is a reasonable estimate of the accuracy of the present calculations of the relativistic energy shifts for Al II, Ca I and Sr I, 15\% for In II, 25\% for Yb I and 20\% for Hg I and Tl II.
405:
406: In Table \ref{tab:final} we present final values of the relativistic energy
407: shifts together with the experimental energies.
408:
409:
410: \section*{Acknowledgments}
411: This work is supported by the Australian Research council.
412:
413: \bibliography{qc}
414:
415: \end{document}
416: