1: %\documentclass[pra,twocolumn,showpacs]{revtex4}
2: \documentclass[pra,preprint,tightenlines,a4paper,showpacs]{revtex4}
3: %\documentclass[pra,preprint,showpacs]{revtex4}
4: %\documentclass[pra,preprint,tightenlines,showpacs]{revtex4}
5: \usepackage{bm,dcolumn,graphicx}
6: %------------------------------------------------------------------
7: \begin{document}
8: %####################################################################
9: \title{High-precision calculations of electric-dipole amplitudes
10: for transitions between low-lying levels of Mg, Ca, and Sr}
11: \author{S.~G.~Porsev}
12: \email{porsev@thd.pnpi.spb.ru},
13: \author{M.~G.~Kozlov}
14: \author{Yu.~G.~Rakhlina}
15: \affiliation{Petersburg Nuclear Physics Institute,
16: Gatchina, Leningrad district, 188300, Russia}
17: \author{A.~Derevianko}
18: \affiliation{Physics Department, University of Nevada, Reno,
19: Nevada 89557}
20: \date{\today}
21:
22: \begin{abstract}
23: To support efforts on cooling and trapping of alkaline-earth
24: atoms and designs of atomic clocks, we performed \emph{ab initio}
25: relativistic many-body calculations of electric-dipole transition
26: amplitudes between low-lying states of Mg, Ca, and Sr. In
27: particular, we report amplitudes for $^1P^o_1 \rightarrow
28: {}^1S_0,{}^3S_1,{}^1D_2$, for $^3P^o_1 \rightarrow
29: {}^1S_0,{}^1D_2$, and for $^3P^o_2 \rightarrow {}^1D_2$
30: transitions. For Ca, the reduced matrix element $\langle 4s4p
31: \,^1\! P^o_1||D||4s^2 \,^1\! S_0 \rangle$ is in a good agreement
32: with a high-precision experimental value deduced from
33: photoassociation spectroscopy [Zinner {\em et al.}, Phys.\ Rev.\
34: Lett.\ {\bf 85}, 2292 (2000) ]. An estimated uncertainty of the
35: calculated lifetime of the $3s3p\,^1\! P^o_1$ state of Mg is a
36: factor of three smaller than that of the most accurate experiment.
37: Calculated binding energies reproduce experimental values within
38: 0.1-0.2\%.
39: \end{abstract}
40:
41: \pacs{31.10.+z, 31.15.Ar, 31.15.Md, 32.70.Cs}
42:
43: \maketitle
44: %####################################################################
45:
46: %=====================
47: \section{Introduction}
48: %=====================
49: Many-body methods have proven to be a highly accurate tool for
50: determination of atomic properties, especially for systems with
51: one valence electron outside a closed-shell core~\cite{Sap98}. For
52: alkali-metal atoms a comparison of highly-accurate experimental
53: data with calculations~\cite{SafJohDer99} allows one to draw a
54: conclusion that modern \emph{ab initio} methods are capable of
55: predicting basic properties of low-lying states with a precision
56: better than 1\%.
57:
58: For {\em divalent} atoms such a comprehensive comparison was
59: previously hindered by a lack of high-precision measurements of
60: radiative lifetimes. Despite the lifetimes of the lowest
61: $nsnp\,^1\!P^o_1$ and $nsnp\,^3\!P^o_1$ states were repeatedly
62: obtained both experimentally and
63: theoretically \cite{Zin,Seng,L80,FF,JonFisGod99,VaeGodHan88,KM80,BFV,WGTG,
64: Lund,PRT,Smith,Hans, Hunter,God,HusR,HusS,Kwong,Droz,Mitch,Whit,
65: Kell}, persistent discrepancies remain. Only very recently, Zinner
66: {\em et al.}~\cite{Zin} have achieved 0.4\% accuracy for the rate
67: of $4s4p\,^1\!P^o_1 \rightarrow 4s^2\,^1\!S_0$ transition in
68: calcium. This high-precision value was deduced from
69: photoassociation spectroscopy of ultracold calcium atoms. One of
70: the purposes of the present work is to test the quality of
71: many-body techniques for two-valence electron systems by comparing
72: our result with the experimental value from Ref.~\cite{Zin}.
73:
74: We extend the earlier work~\cite{PorKozRah00} and report results
75: of relativistic many-body calculation of energy levels and
76: electric-dipole transition amplitudes for Mg, Ca and Sr. The
77: calculations are performed in the framework of
78: configuration-interaction approach coupled with many-body
79: perturbation theory~\cite{DFK,DKPF}. We tabulate
80: electric-dipole amplitudes for $^1P^o_1 \rightarrow
81: {}^1S_0,{}^3S_1,{}^1D_2$, for $^3P^o_1 \rightarrow
82: {}^1S_0,{}^1D_2$, and for $^3P^o_2 \rightarrow {}^1D_2$
83: transitions and estimate theoretical uncertainties.
84:
85: Cooling and trapping experiments with alkaline-earth atoms were
86: recently reported for Mg~\cite{Seng}, Ca~\cite{Zin,Kuro}, and
87: Sr~\cite{Din,Kat}. The prospects of achieving Bose-Einstein
88: condensation were also discussed~\cite{Mg,Zin}. Our accurate
89: transition amplitudes will be helpful in designs of cooling
90: schemes and atomic clocks. In addition, these amplitudes will aid
91: in determination of long-range atomic interactions, required in
92: calculation of scattering lengths and interpretation of
93: cold-collision data. For example, dispersion (van der Waals)
94: coefficient $C_6$ characterizes the leading dipole-dipole
95: interaction of two ground-state atoms at large internuclear
96: separations~\cite{DalDav66}. The coefficient $C_6$ is expressed
97: in terms of energy separations and electric-dipole matrix elements
98: between the ground and excited atomic states. Approximately 80\%
99: of the total value of $C_6$ arises from the principal transition
100: $nsnp\,^1\!P^o_1 - ns^2\,^1\!S_0$, requiring accurate predictions
101: for the relevant matrix element. Therefore our results will be
102: also useful in determination of dispersion coefficients.
103:
104: %===============================
105: \section{Method of calculations}
106: %===============================
107:
108: In atomic-structure calculations, correlations are conventionally
109: separated into three classes: valence-valence, core-valence, and
110: core-core correlations. A strong repulsion of valence electrons
111: has to be treated non-perturbatively, while it is impractical to
112: handle the two other classes of correlations with non-perturbative
113: techniques, such as configuration-interaction (CI) method.
114: Therefore, it is natural to combine many-body perturbation theory
115: (MBPT) with one of the non-perturbative methods. It was
116: suggested \cite{DFK} to use MBPT to construct an effective
117: Hamiltonian $H_{\rm eff}$ defined in the model space of valence
118: electrons. Energies and wavefunctions of low-lying states are
119: subsequently determined using CI approach, i.e. diagonalizing
120: $H_{\rm eff}$ in the valence subspace. Atomic observables are
121: calculated with effective operators~\cite{DKPF}. Following the
122: earlier work, we refer to this method as CI+MBPT formalism.
123:
124: In the CI+MBPT approach the energies and wavefunctions are
125: determined from the Schr\"odinger equation
126: \begin{equation} \label{Eqn_Sh}
127: H_{\rm eff}(E_n) \, | \Phi_n \rangle = E_n \, |\Phi_n \rangle \, ,
128: \end{equation}
129: where the effective Hamiltonian is defined as
130: \begin{equation}\label{Eqn_Heff}
131: H_{\rm eff}(E) = H_{\rm FC} + \Sigma(E).
132: \end{equation}
133: Here $H_{\rm FC}$ is the two-electron Hamiltonian in the frozen
134: core approximation and $\Sigma$ is the
135: energy-dependent correction, involving core excitations. The
136: operator $\Sigma$ completely accounts for the second order of
137: perturbation theory. Determination of the second order corrections
138: requires calculation of one-- and two--electron diagrams. The
139: one--electron diagrams describe an attraction of a valence
140: electron by a (self-)induced core polarization. The two-electron
141: diagrams are specific for atoms with several valence electrons and
142: represent an interaction of a valence electron with core
143: polarization induced by another valence electron.
144:
145: Already at the second order the number of the two--electron
146: diagrams is large and their computation is very time-consuming. In
147: the higher orders the calculation of two-electron diagrams becomes
148: impractical. Therefore we account for the higher orders of MBPT
149: indirectly. It was demonstrated~\cite{Opt} that a proper
150: approximation for the effective Hamiltonian can substantially
151: improve an agreement between calculated and experimental spectra
152: of multielectron atom. One can introduce an energy shift $\delta$
153: and replace $\Sigma(E) \rightarrow \Sigma(E-\delta)$ in the
154: effective Hamiltonian, Eq.~(\ref{Eqn_Heff}).
155: The choice $\delta$=0 corresponds to the Brillouin-Wigner variant
156: of MBPT and the Rayleigh-Schr\"odinger variant is recovered
157: setting $\delta = E_n - E_n^{(0)}$, where $E_n^{(0)}$ is the
158: zero-order energy of level $n$. The latter is more adequate for
159: multielectron systems~\cite{Thoul}; for few-electron systems an
160: intermediate value of $\delta$ is optimal. We have determined
161: $\delta$ from a fit of theoretical energy levels to experimental
162: spectrum. Such an optimized effective Hamiltonian was used in
163: calculations of transition amplitudes.
164:
165: To obtain an effective electric-dipole operator we solved
166: random-phase approximation (RPA) equations, thus summing a certain
167: sequence of many-body diagrams to all orders of MBPT. The RPA
168: describes a shielding of externally applied field by core
169: electrons. We further incorporated one- and two-electron
170: corrections to the RPA to account for a difference between the
171: V$^N$ and V$^{N-2}$ potentials and for the Pauli exclusion
172: principle. In addition, the effective operator included
173: corrections for normalization and structural
174: radiation~\cite{DKPF}. The RPA equations depend on transition
175: frequency and should be solved independently for each transition.
176: However, the frequency dependence was found to be rather weak and
177: we solved these equations only at some characteristic frequencies.
178: To monitor a consistency of the calculations we employed both
179: length (L) and velocity (V) gauges for the electric-dipole operator.
180:
181: The computational procedure is similar to calculations of
182: hyperfine structure constants and electric-dipole amplitudes for
183: atomic ytterbium \cite{PRK1,PRK2}. We consider Mg, Ca and Sr as
184: atoms with two valence electrons above closed cores
185: [1$s$,...,2$p^6$], [1$s$,...,3$p^6$], and [1$s$,...,4$p^6$],
186: respectively \cite{sidenote}.
187: One-electron basis set for Mg included 1$s$--13$s$, 2$p$--13$p$,
188: 3$d$--12$d$, and 4$f$--11$f$ orbitals, where the core- and 3,4$s$,
189: 3,4$p$, 3,4$d$, and 4$f$ orbitals were Dirac-Hartree-Fock (DHF)
190: ones, while all the rest were virtual orbitals. The orbitals
191: 1$s$--3$s$ were constructed by solving the DHF equations in V$^N$
192: approximation, 3$p$ orbitals were obtained in the V$^{N-1}$
193: approximation, and 4$s$, 4$p$, 3,4$d$, and 4$f$ orbitals were
194: constructed in the V$^{N-2}$ approximation. We determined virtual
195: orbitals using a recurrent procedure, similar to
196: Ref.~\cite{Bogdan} and employed in previous
197: work~\cite{DFK,DKPF,PRK1,PRK2}. The one-electron basis set for Ca
198: included 1$s$--13$s$, 2$p$--13$p$, 3$d$--12$d$, and 4$f$--11$f$
199: orbitals, where the core- and 4$s$, 4$p$, and 3$d$ orbitals are
200: DHF ones, while the remaining orbitals are the virtual orbitals.
201: The orbitals 1$s$--4$s$ were constructed by solving the DHF
202: equations in the V$^N$ approximation, and 4$p$ and 3$d$ orbitals
203: were obtained in the V$^{N-1}$ approximation. Finally, the
204: one-electron basis set for Sr included 1$s$--14$s$, 2$p$--14$p$,
205: 3$d$--13$d$, and 4$f$--13$f$ orbitals, where the core- and 5$s$,
206: 5$p$, and 4$d$ orbitals are DHF ones, and all the rest are the
207: virtual orbitals. The orbitals 1$s$--5$s$ were constructed by
208: solving the DHF equations in the V$^N$ approximation, and 5$p$ and
209: 4$d$ orbitals were obtained in the V$^{N-1}$ approximation.
210: Configuration-interaction states were formed using these
211: one-particle basis sets. It is worth emphasizing that the employed
212: basis sets were sufficiently large to obtain numerically converged
213: CI results. A numerical solution of random-phase approximation
214: equations required an increase in the number of virtual
215: orbitals. Such extended basis sets included $1s$--$ks$,
216: $2p$--$kp$, $3d$--($k$-1)$d$, $4f$--($k$-4)$f$, and
217: $5g$--($k$-8)$g$ orbitals, where $k$=19,20,21 for Mg, Ca, and Sr,
218: respectively. Excitations from all core shells were included in
219: the RPA setup.
220:
221: %===============================
222: \section{Results and discussion}
223: %===============================
224: \subsection{Energy levels}
225: In Tables~\ref{Mg_E} -- \ref{Sr_E} we present calculated energies
226: of low-lying states for Mg, Ca, and Sr and compare them with
227: experimental values. The two-electron binding energies were
228: obtained both in the framework of conventional
229: configuration-interaction method and using the formalism of
230: CI coupled with many-body perturbation theory. Already
231: at the CI stage the agreement of the calculated and experimental
232: energies is at the level of 5\%. The addition of many-body
233: corrections to the Hamiltonian improves the accuracy by
234: approximately an order of magnitude. Finally, with an optimal
235: choice of parameter $\delta$ the agreement with experimental
236: values improves to 0.1--0.2\%.
237:
238: Compared to the binding energies, fine-structure splitting of
239: triplet states and singlet-triplet energy differences represent a
240: more stringent test of our method. For the $^3P^o_{1,2,3}$-states
241: the fine-structure splitting is reproduced with an accuracy of
242: several per cent in the pure CI for all the three atoms, while the
243: $^3P^o_1$ -- $^1P^o_1$ energy differences are less accurate
244: (especially for Ca and Sr). As demonstrated in Ref.~\cite{Opt}, the
245: two-electron exchange Coulomb integral $R_{np,ns,ns,np}$
246: ($n$=3,4,5 for Mg, Ca, and Sr, respectively) determining the
247: splitting between $^3P^o_1$ and $^1P^o_1$ states is very sensitive
248: to many-body corrections. Indeed, with these corrections included,
249: the agreement with the experimental data improves to 1-2\% for all
250: the three atoms.
251:
252: The case of the even-parity $^{3,1}D_J$-states is even more
253: challenging. For Ca, these four states are practically degenerate
254: at the CI stage. A repulsion of the level $^1D_2$ from the
255: upper-lying levels of $np^2$ configuration pushes it down to the
256: level $^3D_2$ and causes their strong mixing. As seen from
257: Table~\ref{Ca_E} these states are separated only by 10 cm$^{-1}$,
258: while the experimental energy difference is 1550 cm$^{-1}$. As a
259: result, an accurate description of superposition of $^3D_2$ and
260: $^1D_2$ states is important. The $^3D_2$ -- $^1D_2$ splitting is
261: restored when the many-body corrections are included in the
262: effective Hamiltonian. These corrections properly account for core
263: polarization screening an interaction between $sd$ and $p^2$
264: configurations.
265:
266: For Sr, the fine-structure splittings of $^3D_J$ states and energy
267: difference between the $^3D_J$ and the $^1D_2$ levels are also
268: strongly underestimated in the pure CI method. Again the inclusion
269: of the many-body corrections substantially improves the splittings
270: between the $D$-states. It is worth emphasizing, that for such an
271: accurate analysis a number of effects was taken into account,
272: i.e., spin-orbit interaction, configuration interaction, and
273: core-valence correlations. A proper account for all these effects
274: is of particular importance for determination of electric-dipole
275: amplitudes forbidden in $LS$-coupling, such as for $^3P^o_J
276: \rightarrow {}^1S_0,{}^1D_2$ transitions.
277:
278: \subsection{Transition amplitudes}
279:
280: In this section we present calculations of electric-dipole (E1)
281: amplitudes for $^{3,1}P^o_1 \rightarrow {}^1S_0$, $^{3,1}P^o_1
282: \rightarrow {}^1D_2$, $^3P^o_2 \rightarrow {}^1D_2$, and $^1P^o_1
283: \rightarrow {}^3S_1$ transitions. The calculated reduced matrix
284: elements for Mg, Ca, and Sr are presented in
285: Tables~\ref{Tab_E1_allowed} and \ref{Tab_E1_inter}. For a
286: transition $I \rightarrow F$ the Einstein rate coefficients for
287: spontaneous emission (in $1/s$) are expressed in terms of these
288: reduced matrix elements $\langle F || D || I \rangle$ (a.u.) and
289: wavelengths $\lambda$ (\AA) as
290: \begin{equation}
291: A_{FI} = \frac{2.02613 \times 10^{18}} {\lambda^3 }
292: \frac{ |\langle F || D || I \rangle|^2}{ 2 J_I +1} \, .
293: \end{equation}
294: A number of long-range atom-atom interaction coefficients could
295: be directly obtained from the calculated matrix elements. At large
296: internuclear separations $R$ an atom in a state $|A \rangle$
297: predominantly interacts with a like atom in a state $|B\rangle$
298: through a potential $V(R) \approx \pm C_3/R^3$, provided an
299: electric-dipole transition between the two atomic states $|A
300: \rangle$ and $|B \rangle$ is allowed. The coefficient $C_3$ is
301: given by
302: \begin{equation}
303: |C_3| = |\langle A||D||B\rangle|^{2} \, \sum_{\mu=-1}%
304: ^{1}\left( 1+\delta_{\mu,0}\right) \left(
305: \begin{tabular}
306: [c]{ccc}%
307: $J_{A}$ & $1$ & $J_{B}$\\
308: $-\frac{\Omega+\mu}{2}$ & $\mu$ & $\frac{\Omega-\mu}{2}$%
309: \end{tabular}
310: \right) ^{2}%
311: \, ,
312: \end{equation}
313: where $\Omega$ is the conventionally defined sum of projections of
314: total angular momenta on internuclear axis.
315:
316: From a solution of the eigen-value problem, Eq.~(\ref{Eqn_Sh}), we
317: obtained wave functions, constructed effective dipole operators,
318: and determined the transition amplitudes. The calculations were
319: performed within both traditional configuration-interaction
320: method and CI coupled with the many-body perturbation theory.
321: The comparison of the CI and the CI+MBPT values allows
322: us to estimate an accuracy of our calculations. As it was
323: mentioned above, to monitor the consistency of the calculations, we
324: determined the amplitudes using both length and velocity gauges
325: for the dipole operator. In general, dipole amplitudes calculated in
326: the velocity gauge are more sensitive to many-body corrections; we
327: employ the length form of the dipole operator in our final tabulation.
328:
329: We start the discussion with the amplitudes for the principal
330: $nsnp\,^1\!P^o_1 \rightarrow ns^2\, ^1\!S_0$ transitions ($n=3$
331: for Mg, $n=4$ for Ca, and $n=5$ for Sr). Examination of
332: Table~\ref{Tab_E1_allowed} reveals that the many-body effects
333: reduce the L-gauge amplitudes by 1.6\% for Mg, 5.5\% for Ca,
334: and 6.4\% for Sr. Further, the MBPT corrections bring the length
335: and velocity-form results into a closer agreement. For example,
336: for Sr at the CI level the velocity and length forms differ by
337: 2.7\% and this discrepancy is reduced to 0.8\% in the CI+MBPT
338: calculations.
339:
340: A dominant theoretical uncertainty of the employed CI+MBPT method
341: is due to impossibility to account for all the orders of many-body
342: perturbation theory. It is worth emphasizing that in our CI
343: calculations the basis sets were saturated and the associated
344: numerical errors were negligible. We expect that the theoretical
345: uncertainty is proportional to the determined many-body
346: correction. In addition, we take into account the proximity of the
347: amplitudes obtained in the L- and V-gauges. We estimate the
348: uncertainties for the $nsnp\,^1\!P^o_1 \rightarrow ns^2\, ^1\!S_0$
349: transition amplitudes as 25--30\% of the many-body corrections in
350: the length gauge. The final values for $\langle nsnp\,^1\!P^o_1
351: ||D|| ns^2\, ^1\!S_0 \rangle$, recommended from the present work,
352: are 4.03(2) for Mg, 4.91(7) for Ca, and 5.28(9) a.u. for Sr.
353:
354: We present a comparison of our results for $\langle
355: nsnp\,^1\!P^o_1 ||D|| ns^2\, ^1\!S_0 \rangle$ with experimental
356: data in Table~\ref{Tab_E1_allowed} and in Fig.~\ref{Fig_princ}.
357: Our estimated accuracy for Mg is a factor of three better than
358: that of the most accurate experiment and for Sr is comparable to
359: experimental precision. For Ca, the dipole matrix element of the
360: $^1\!P^o_1 \rightarrow {}^1\!S_0$ was recently determined with a
361: precision of 0.2\% by Zinner {\em et al.} \cite{Zin} using
362: photoassociation spectroscopy of ultracold Ca atoms. While our
363: result is in harmony with their value, the experimental accuracy
364: is substantially better. An updated analysis~\cite{Tiemann} of
365: photoassociation spectra of Ref.~\cite{Zin} leads to a somewhat
366: better agreement with our calculated value.
367:
368: A very extensive compilation of earlier theoretical results for
369: the $ ^1\!P^o_1 \rightarrow {}^1\!S_0$ transition amplitudes can be
370: found in Ref.~\cite{FF} for Mg and in Ref.~\cite{BFV} for Ca. In a
371: very recent multiconfiguration Hartree-Fock (MCHF) calculations
372: for Mg~\cite{JonFisGod99} the authors have determined $ \langle
373: 3s3p\,^1\!P^o_1||D|| 3s^2\,^1\!S_0 \rangle =4.008 $ a.u. This value
374: agrees with our final result of 4.03(2) a.u. For heavier Sr the
375: correlation effects are especially pronounced and only a few
376: calculations were performed. For example, MCHF calculations for
377: Sr~\cite{VaeGodHan88} found in the length gauge $ \langle
378: 5s5p\,^1\!P^o_1||D|| 5s^2\,^1\!S_0 \rangle = 5.67$ a.u. By contrast
379: to the present work, the core-polarization effects were not
380: included in this analysis. As a result, this calculated value is
381: in a good agreement with our result 5.63 a.u. obtained at the CI stage,
382: but differs from the final value $5.28(9)$ a.u.
383:
384: Another nonrelativistically allowed transition is $^1P^o_1
385: \rightarrow {}^1D_2$ and one could expect that this amplitude can
386: be determined with a good accuracy. For Mg this is really so.
387: However, for Ca and Sr an admixture of the configuration $p^2$
388: brings about large corrections to this amplitude, especially in
389: the velocity gauge. Another complication is the following. The
390: matrix element of electric-dipole operator can be represented in
391: the V-gauge as (atomic units $\hbar = |e| = m_e = 1$ are used):
392: \begin{equation}
393: \langle F |{\bf D}| I \rangle = i \,c \, \langle F|{\bm{\alpha}}|
394: I \rangle /(E_I - E_F).
395: \end{equation}
396: Here $c$ is the speed of light, $E_I$ and $E_F$ are the energies
397: of initial and final states, and $\bm{\alpha}$ are the Dirac
398: matrices. For the $^1P^o_1 \rightarrow {}^1D_2$ transition in Ca
399: and Sr the energy denominator is approximately 0.01 a.u. Because
400: the E1-amplitudes of these transitions $\sim 1$ a.u. (see
401: Table~\ref{Tab_E1_allowed}), the respective numerators are of the
402: order of 0.01 a.u. Correspondingly the matrix elements $\langle
403: F|{\bm{\alpha}}|I\rangle$ are small and are very sensitive to
404: corrections, i.e., the V-gauge results are unstable. As a result
405: we present only the L-gauge values for $^1P^o_1 \rightarrow
406: {}^1D_2$ E1 amplitudes for Ca and Sr. An absence of reliable
407: results in V-gauge hampers an estimate of the accuracy, so we
408: rather conservatively take it to be 25\%. Note that even with such
409: a large uncertainty our value for Sr significantly differs from
410: the experimental value \cite{Hunter}. The measurement in
411: \cite{Hunter} has been carried out on the $^1D_2 \rightarrow
412: {}^1S_0$ transition and an interference between
413: electric-quadrupole (E2) and Stark-induced dipole amplitudes was
414: observed. In order to determine the transition rate a theoretical
415: value of the E2-amplitude for the $^1D_2 \rightarrow {}^1S_0$
416: transition was taken from \cite{Bausch}. It may be beneficial
417: either to measure directly the rate of the E1-transition $^1P^o_1
418: \rightarrow {}^1D_2$ or to measure the rate of the E2-transition
419: $^1D_2 \rightarrow {}^1S_0$.
420:
421: For the $^3P^o_J \rightarrow {}^1S_0,{}^1D_2$ transitions the
422: respective E1-amplitudes are small; these are nonrelativistically
423: forbidden intercombination transitions and consequently their
424: amplitudes are proportional to spin-orbit interaction. The
425: calculated reduced matrix elements are presented in
426: Table~\ref{Tab_E1_inter}.
427:
428: One can see from Tables~\ref{Mg_E} --~\ref{Sr_E} that the MBPT
429: corrections to the fine structure splittings are large, amplifying
430: significance of higher order many-body corrections. In addition,
431: higher order corrections in the fine-structure constant $\alpha$
432: to the Dirac-Coulomb Hamiltonian are also important here. As
433: demonstrated in Ref.~\cite{FF}, the Breit interaction reduces the
434: dipole amplitude of $^3P^o_1 \rightarrow {}^1S_0$ transition in Mg
435: by 5\%. At the same time for all the intercombination transitions
436: the agreement between L- and V-gauges is at the level of 6-8\%. We
437: conservatively estimate the uncertainties of the calculated
438: intercombination E1 amplitudes to be 10--12\%.
439:
440: To reiterate, we carried out calculations of energies of low-lying
441: levels and electric-dipole amplitudes between them for divalent
442: atoms Mg, Ca, and Sr. We employed {\em ab initio} relativistic
443: configuration interaction method coupled with many-body
444: perturbation theory. The calculated removal energies reproduce
445: experimental values within 0.1-0.2\%. A special emphasis has been
446: put on accurate determination of electric-dipole amplitudes for
447: principal transitions $nsnp\,^1\!P^o_1 \rightarrow ns^2\,^1\!S_0$.
448: For these transitions, we estimated theoretical uncertainty to be
449: 0.5\% for Mg, 1.4\% for Ca, and 1.7\% for Sr. For Ca, the reduced
450: matrix element $\langle 4s4p \,^1\! P^o_1||D||4s^2 \,^1\! S_0
451: \rangle$ is in a good agreement with a high-precision experimental
452: value~\cite{Zin}. An estimated uncertainty of the calculated
453: lifetime of the lowest $^1\! P^o_1$ state for Mg is a factor of
454: three smaller than that of the most accurate experiment. In
455: addition, we evaluated electric-dipole amplitudes and estimated
456: theoretical uncertainties for $^1P^o_1 \rightarrow
457: {}^3\!S_1,{}^1\!D_2$, $^3\!P^o_1 \rightarrow {}^1\!S_0,{}^1D_2$,
458: and for $^3\!P^o_2 \rightarrow {}^1\!D_2$ transitions. Our results
459: could be useful in designs of cooling schemes and atomic clocks,
460: and for accurate description of long-range atom-atom interactions
461: needed in interpretation of cold-collision data.
462:
463: \acknowledgments We would like to thank H. Katori, C. Oates, and
464: F. Riehle for stimulating discussions. This work was supported in
465: part by the Russian Foundation for Basic Researches (grant No
466: 98-02-17663). The work of A.D. was partially supported by the
467: Chemical Sciences, Geosciences and Biosciences Division of the
468: Office of Basic Energy Sciences, Office of Science, U.S.
469: Department of Energy.
470:
471: %####################################################################
472: \begin{table}
473: \caption{Two-electron binding energies E$_{\rm val}$ in a.u. and
474: energy differences $\Delta$ (cm$^{-1}$) for low-lying levels of
475: Mg.}
476: \label{Mg_E}
477: %\begin{ruledtabular}
478: \begin{tabular}{lcdcdcdc}
479: \hline \hline
480: & & \multicolumn{2}{ c}{\qquad CI}
481: & \multicolumn{2}{ c}{\qquad CI+MBPT}
482: & \multicolumn{2}{c}{\qquad Experiment} \\
483: Config.& Level
484: &\multicolumn{1}{c}{\qquad E$_{\rm val}$}
485: &\multicolumn{1}{c}{$\Delta$}
486: &\multicolumn{1}{c}{\qquad E$_{\rm val}$}
487: &\multicolumn{1}{c}{$\Delta$}
488: &\multicolumn{1}{c}{\qquad E$_{\rm val}$}
489: &\multicolumn{1}{c}{$\Delta$} \\
490: \hline
491: $3s^2$ & $^1S_0$ & 0.819907 & ---
492: & 0.833556 & --- & 0.833518 \footnotemark[1] & --- \\
493: $3s\,4s$ & $^3S_1$ & 0.635351 & 40505
494: & 0.645853 & 41196 & 0.645809 & 41197.4 \\
495: $3s\,4s$ & $^1S_0$ & 0.624990 & 42779
496: & 0.635283 & 43516 & 0.635303 & 43503.1 \\
497: $3s\,3d$ & $^1D_2$ & 0.613603 & 45278
498: & 0.621830 & 46469 & 0.622090 & 46403.1 \\
499: \hline
500: $3s\,3p$ & $^3P^o_0$ & 0.724170 & 21012
501: & 0.733896 & 21879 & 0.733961 & 21850.4 \\
502: $3s\,3p$ & $^3P^o_1$ & 0.724077 & 21032
503: & 0.733796 & 21901 & 0.733869 & 21870.4 \\
504: $3s\,3p$ & $^3P^o_2$ & 0.723889 & 21073
505: & 0.733596 & 21945 & 0.733684 & 21911.1 \\
506: $3s\,3p$ & $^1P^o_1$ & 0.662255 & 34601
507: & 0.674226 & 34975 & 0.673813 & 35051.4 \\
508: \hline \hline
509: \end{tabular}
510: %\end{ruledtabular}
511: \footnotemark[1]{Two~electron binding energy of the ground
512: state is determined as a sum of the first two ionization
513: potentials IP~(Mg) and IP~(Mg$^+$), where IP~(Mg) = 61669.1
514: cm$^{-1}$ and IP~(Mg$^+$)$ = 121267.4$ cm$^{-1}$ \cite{Moore}}.
515: \end{table}
516: %####################################################################
517: \begin{table}
518: \caption{Two-electron binding energies in a.u. and energy
519: differences $\Delta$ in cm$^{-1}$ for the low-lying levels of
520: Ca.}
521: \label{Ca_E}
522: %\begin{ruledtabular}
523: \begin{tabular}{lcdcdcdc}
524: \hline \hline
525: & & \multicolumn{2}{ c}{\qquad CI \footnotemark[1]}
526: & \multicolumn{2}{ c}{\qquad CI+MBPT}
527: & \multicolumn{2}{c}{\qquad Experiment} \\
528: Config.& Level
529: &\multicolumn{1}{c}{\qquad E$_{\rm val}$}
530: &\multicolumn{1}{c}{$\Delta$}
531: &\multicolumn{1}{c}{\qquad E$_{\rm val}$}
532: &\multicolumn{1}{c}{$\Delta$}
533: &\multicolumn{1}{c}{\qquad E$_{\rm val}$}
534: &\multicolumn{1}{c}{$\Delta$} \\
535: \hline
536: $4s^2$ & $^1S_0$ & 0.636590 & ---
537: & 0.661274 & --- & 0.660927 \footnotemark[2] & --- \\
538: $4s\,3d$ & $^3D_1$ & 0.528838 & 23649
539: & 0.567744 & 20527 & 0.568273 & 20335.3 \\
540: $4s\,3d$ & $^3D_2$ & 0.528868 & 23642
541: & 0.567656 & 20547 & 0.568209 & 20349.2 \\
542: $4s\,3d$ & $^3D_3$ & 0.528820 & 23653
543: & 0.567517 & 20577 & 0.568110 & 20371.0 \\
544: $4s\,3d$ & $^1D_2$ & 0.528824 & 23652
545: & 0.559734 & 22285 & 0.561373 & 21849.6 \\
546: $4s\,5s$ & $^3S_1$ & 0.498205 & 30372
547: & 0.517490 & 31557 & 0.517223 & 31539.5 \\
548: \hline
549: $4s\,4p$ & $^3P^o_0$ & 0.574168 & 13700
550: & 0.591521 & 15309 & 0.591863 & 15157.9 \\
551: $4s\,4p$ & $^3P^o_1$ & 0.573942 & 13750
552: & 0.591274 & 15363 & 0.591625 & 15210.1 \\
553: $4s\,4p$ & $^3P^o_2$ & 0.573486 & 13850
554: & 0.590774 & 15473 & 0.591143 & 15315.9 \\
555: $4s\,4p$ & $^1P^o_1$ & 0.530834 & 23211
556: & 0.553498 & 23654 & 0.553159 & 23652.3 \\
557: \hline \hline
558: \end{tabular}
559: %\end{ruledtabular}
560: \small \footnotemark[1]{Note~that~the conventional CI fails to
561: recover the correct ordering of $D$-states.}
562: \footnotemark[2]{For the ground state E$_{\rm val}$=
563: IP~(Ca)+IP~(Ca$^+$), where IP~(Ca) = 49304.8 cm$^{-1}$ and IP~(Ca$^+$)
564: = 95752.2 cm$^{-1}$ \cite{Moore}}.
565: \end{table}
566: %####################################################################
567: \begin{table}
568: \caption{Two-electron binding energies in a.u. and energy
569: differences $\Delta$ in cm$^{-1}$ for the low-lying levels of Sr.}
570: \label{Sr_E}
571: %\begin{ruledtabular}
572: \begin{tabular}{lcdcdcdc}
573: \hline \hline
574: & & \multicolumn{2}{ c}{\qquad CI}
575: & \multicolumn{2}{ c}{\qquad CI+MBPT}
576: & \multicolumn{2}{c}{\qquad Experiment} \\
577: Config.& Level
578: &\multicolumn{1}{c}{\qquad E$_{\rm val}$}
579: &\multicolumn{1}{c}{$\Delta$}
580: &\multicolumn{1}{c}{\qquad E$_{\rm val}$}
581: &\multicolumn{1}{c}{$\Delta$}
582: &\multicolumn{1}{c}{\qquad E$_{\rm val}$}
583: &\multicolumn{1}{c}{$\Delta$} \\
584: \hline
585: $5s^2$ & $^1S_0$ & 0.586538 & ---
586: & 0.614409 & --- & 0.614601 \footnotemark[1] & --- \\
587: $5s\,4d$ & $^3D_1$ & 0.497148 & 19619
588: & 0.532110 & 18063 & 0.531862 & 18159.1 \\
589: $5s\,4d$ & $^3D_2$ & 0.497077 & 19635
590: & 0.531809 & 18129 & 0.531590 & 18218.8 \\
591: $5s\,4d$ & $^3D_3$ & 0.496941 & 19664
592: & 0.531298 & 18242 & 0.531132 & 18319.3 \\
593: $5s\,4d$ & $^1D_2$ & 0.494339 & 20235
594: & 0.522311 & 20213 & 0.522792 & 20149.7 \\
595: $5s\,6s$ & $^3S_1$ & 0.460940 & 27566
596: & 0.481533 & 29162 & 0.482291 & 29038.8 \\
597: \hline
598: $5s\,5p$ & $^3P^o_0$ & 0.529636 & 12489
599: & 0.548754 & 14410 & 0.549366 & 14317.5 \\
600: $5s\,5p$ & $^3P^o_1$ & 0.528850 & 12662
601: & 0.547896 & 14598 & 0.548514 & 14504.4 \\
602: $5s\,5p$ & $^3P^o_2$ & 0.527213 & 13021
603: & 0.546079 & 14997 & 0.546718 & 14898.6 \\
604: $5s\,5p$ & $^1P^o_1$ & 0.491616 & 20833
605: & 0.515901 & 21621 & 0.515736 & 21698.5 \\
606: \hline \hline
607: \end{tabular}
608: %\end{ruledtabular}
609: \small \footnotemark[1]{For~the~ground state E$_{\rm val}$ =
610: IP~(Sr)+IP~(Sr$^+$), where IP~(Sr) = 45925.6 cm$^{-1}$ and
611: IP~(Sr$^+$) = 88964.0 cm$^{-1}$ \cite{Moore}}.
612: \end{table}
613:
614: %####################################################################
615: \begin{table}
616: \caption{ Reduced electric-dipole matrix elements for transitions
617: allowed in $LS$-coupling. $n$ is the principal quantum number of
618: the first valence $s$ and $p$ shells and $m$ corresponds to the
619: first valence $d$ shell; $n=3$ for Mg, 4 for Ca, and 5 for Sr;
620: $m=3$ for Mg and Ca, and 4 for Sr. All values are in a.u.}
621: \label{Tab_E1_allowed}
622: %\begin{ruledtabular}
623: \begin{tabular}{ldddddd}
624: \hline \hline
625: &\multicolumn{2}{c}{Mg} & \multicolumn{2}{c}{Ca} &
626: \multicolumn{2}{c}{Sr}
627: \\
628: &\multicolumn{1}{c}{CI} & \multicolumn{1}{c}{CI+MBPT} &
629: \multicolumn{1}{c}{CI} & \multicolumn{1}{c}{CI+MBPT} &
630: \multicolumn{1}{c}{CI} & \multicolumn{1}{c}{CI+MBPT}
631: \\
632: \hline \multicolumn{7}{c}{$\langle nsnp
633: \,^1\!P^o_1 ||D|| ns^2\, ^1\!S_0 \rangle$}\\
634: L-gauge & 4.09 & 4.03 & 5.20 & 4.91 & 5.63 & 5.28 \\
635: V-gauge & 4.06 & 4.04 & 5.11 & 4.93 & 5.48 & 5.32 \\
636: Final value & \multicolumn{2}{d }{4.03(2)}
637: & \multicolumn{2}{d }{4.91(7)}
638: & \multicolumn{2}{d}{5.28(9)} \\
639: Experiment & \multicolumn{2}{d }{4.15(10) \footnotemark[1]}
640: & \multicolumn{2}{d }{4.967(9) \footnotemark[2]}
641: & \multicolumn{2}{d }{5.57(6) \footnotemark[3]}
642: \\ & \multicolumn{2}{d }{4.06(10) \footnotemark[4]}
643: & \multicolumn{2}{d }{4.99(4) \footnotemark[3]}
644: & \multicolumn{2}{d }{5.40(8) \footnotemark[5]}
645: \\ & \multicolumn{2}{d }{4.12(6) \footnotemark[6]}
646: & \multicolumn{2}{d }{4.93(11) \footnotemark[7]}\\
647: \hline \multicolumn{7}{c}{$\langle nsnp \,
648: ^1\!P^o_1 ||D||nsmd\, {}^1\!D_2\rangle$}\\
649: L-gauge & 4.43 & 4.62 & & 1.16 & 1.75 & 1.92
650: \\
651: V-gauge & 4.47 & 4.59 %&\mbox{---}&\mbox{---}&\mbox{---}&\mbox{---}
652: \\
653: %& 1.62 & 12.9 & 0.26 \\
654: Final value & \multicolumn{2}{d }{4.62(5)}
655: & \multicolumn{2}{d }{1.2(3)}
656: & \multicolumn{2}{d }{1.9(4)}
657: \\
658: Experiment & \multicolumn{2}{d }{}
659: & \multicolumn{2}{d }{}
660: & \multicolumn{2}{d }{1.24(18)\footnotemark[8]} \\
661: % & \multicolumn{2}{c }{}
662: % & \multicolumn{2}{c }{}
663: % & \multicolumn{2}{c}{} \\
664: \hline \hline
665: \end{tabular}
666: %\end{ruledtabular}
667: \footnotemark[1]{Ref.~\cite{L80};}~%
668: \footnotemark[2]{Ref.~\cite{Zin};}~%
669: \footnotemark[3]{Ref.~\cite{KM80};}
670: \footnotemark[4]{Ref.~\cite{Lund};}
671: \footnotemark[5]{Ref.~\cite{PRT};}
672: \footnotemark[6]{Ref.~\cite{Smith};}
673: \footnotemark[7]{Ref.~\cite{Hans};}
674: \footnotemark[8]{Ref.~\cite{Hunter}.}
675: \end{table}
676:
677:
678: %####################################################################
679: \begin{table}
680: \caption{ Reduced electric-dipole matrix elements for {\em
681: intercombination} transitions. $n$ is the principal quantum number
682: of the first valence $s$ and $p$ shells and $m$ corresponds to the
683: first valence $d$ shell; $n=3$ for Mg, 4 for Ca, and 5 for Sr;
684: $m=3$ for Mg and Ca, and 4 for Sr. All values are in a.u.}
685: \label{Tab_E1_inter}
686: %\begin{ruledtabular}
687: \begin{tabular}{ldddddd}
688: \hline \hline
689: &\multicolumn{2}{c}{Mg} & \multicolumn{2}{c}{Ca} &
690: \multicolumn{2}{c}{Sr}
691: \\
692: &\multicolumn{1}{c}{CI} & \multicolumn{1}{c}{CI+MBPT} &
693: \multicolumn{1}{c}{CI} & \multicolumn{1}{c}{CI+MBPT} &
694: \multicolumn{1}{c}{CI} & \multicolumn{1}{c}{CI+MBPT}
695: \\
696: %\hline
697:
698: \hline
699: %
700: \multicolumn{7}{c}{$ \langle nsnp\,^3\!P^o_1||D||ns^2\,^1\!S_0\rangle$}\\
701: L-gauge & 0.0055 & 0.0064 & 0.027 & 0.034 & 0.12 & 0.16 \\
702: V-gauge & 0.0062 & 0.0062 & 0.030 & 0.032 & 0.13 & 0.17 \\
703: Final value & \multicolumn{2}{d }{0.0064(7)}
704: & \multicolumn{2}{d }{0.034(4)}
705: & \multicolumn{2}{d }{0.160(15)} \\
706: Experiment & \multicolumn{2}{d }{0.0053(3) \footnotemark[1]}
707: & \multicolumn{2}{d }{0.0357(4) \footnotemark[2]}
708: & \multicolumn{2}{d }{0.1555(16)\footnotemark[3]}
709: \\ & \multicolumn{2}{d }{0.0056(4) \footnotemark[4]}
710: & \multicolumn{2}{d }{0.0352(10)\footnotemark[5]}
711: & \multicolumn{2}{d }{0.1510(18)\footnotemark[5]}
712: \\ & \multicolumn{2}{d }{0.0061(10)\footnotemark[6]}
713: & \multicolumn{2}{d }{0.0357(16)\footnotemark[7]}
714: & \multicolumn{2}{d }{0.1486(17)\footnotemark[8]} \\
715: \hline
716:
717: \multicolumn{7}{c}{$\langle nsnp\,^1\!P^o_1|| D ||
718: ns(n+1)s\,^3\!S_1\rangle$}\\
719: L-gauge & 0.0088 & 0.0097 & 0.035 & 0.043 & 0.15 & 0.19 \\
720: V-gauge & 0.0089 & 0.0101 & 0.035 & 0.045 & 0.15 & 0.20 \\
721: Final value & \multicolumn{2}{d }{0.0097(10)}
722: & \multicolumn{2}{d }{0.043(5)}
723: & \multicolumn{2}{d }{0.19(2)} \\
724: \hline
725: %
726: %
727: \multicolumn{7}{c}{$\langle nsnp\,^3\!P^o_1||D|| nsmd\,^1\!D_2\rangle$}\\
728: L-gauge & 0.0052 & 0.0049 & & 0.059 & 0.33 & 0.19 \\
729: V-gauge & 0.0050 & 0.0047 & & 0.061 & 0.36 & 0.18 \\
730: Final value & \multicolumn{2}{d }{0.0049(5)}
731: & \multicolumn{2}{d }{0.059(6)}
732: & \multicolumn{2}{d }{0.19(2)} \\
733: \hline
734: %
735: \multicolumn{7}{c}{$\langle nsnp\,^3\!P^o_2||D|| nsmd\,^1\!D_2 \rangle$}\\
736: L-gauge & 0.0039 & 0.0031 & & 0.028 & 0.15 & 0.10 \\
737: V-gauge & 0.0041 & 0.0032 & & 0.024 & 0.16 & 0.06 \\
738: Final value & \multicolumn{2}{d }{0.0031(4)}
739: & \multicolumn{2}{d }{0.028(3)}
740: & \multicolumn{2}{d }{0.10(2)} \\
741: \hline \hline
742: \end{tabular}
743: %\end{ruledtabular}
744: %\footnotemark[8]{}
745: %\footnotemark[9]{Ref.~\cite{Hunter};}
746: \footnotemark[1]{Ref.~\cite{God};}~%
747: \footnotemark[2]{Ref.~\cite{HusR};}~%
748: \footnotemark[3]{Ref.~\cite{HusS};}
749: \footnotemark[4]{Ref.~\cite{Kwong};}
750: \footnotemark[5]{Ref.~\cite{Droz};}
751: \footnotemark[6]{Ref.~\cite{Mitch};}
752: \footnotemark[7]{Ref.~\cite{Whit};}
753: \footnotemark[8]{Ref.~\cite{Kell}.}
754: \end{table}
755:
756: \begin{figure}
757: \centerline{\includegraphics*[scale=0.75]{1p1all.eps}}
758: \caption{ Comparison of calculated reduced matrix elements
759: $\langle nsnp\,^1\!P^o_1 ||D|| ns^2\, ^1\!S_0 \rangle$
760: with experimental data in a.u.
761: \label{Fig_princ} }
762: \end{figure}
763:
764: %####################################################################
765: \bibliographystyle{apsrev}
766: %####################################################################
767: \begin{thebibliography}{99}
768: \bibitem{Sap98}
769: J.~Sapirstein, Rev.\ Mod.\ Phys. \textbf{70}, 55 (1998) and
770: references therein.
771: \bibitem{SafJohDer99} see, for example,
772: M.S.~Safronova, W.R.~Johnson, and A.~Derevianko, Phys.\ Rev.\ A
773: {\bf 60}, 4476 (1999).
774: \bibitem{Zin}
775: G.~Zinner, T.~Binnewies, F.~Riehle, and E.~Tiemann,
776: Phys.~Rev.~Lett. {\bf 85}, 2292 (2000).
777: \bibitem{Seng}
778: K.~Sengstock, U.~Sterr, G.~Hennig, D.~Bettermann, J.H.~Muller, and
779: W.~Ertmer, Opt.~Commun. {\bf 103}, 73 (1993).
780: \bibitem{L80}
781: L.~Liljeby, A.~Lindgard, S.~Mannervik, E.~Veje, and B.~Jelencovic,
782: Phys.~Scr. {\bf 21}, 805 (1980);
783: \bibitem{FF}
784: P.~J\"{o}nsson, and C.~F. Fischer, J.~Phys.~B {\bf 30}, 5861
785: (1997).
786: \bibitem{JonFisGod99} P.~J\"{o}nsson, C.F.~Fischer, and
787: M.R.~Godefroid, J.\ Phys.\ B {\bf 32}, 1233 (1999).
788: \bibitem{VaeGodHan88} N. Vaeck, M. Godefroid, and J.E. Hansen,
789: Phys.\ Rev.\ A {\bf 38} 2830 (1988).
790: \bibitem{BFV}
791: T.~Brage, C.F.~Fischer, N.~Vaeck, M.~Godefroid, and A.~Hibbert,
792: Phys.~Scr. {\bf 48}, 533 (1993) (and references therein).
793: \bibitem{WGTG}
794: H.~G.~C.~Werij, C.~H.~Greene, C.~E.~Theodosiou and A.~Gallagher,
795: Phys.~Rev.~A {\bf 46}, 1248 (1992) (and references therein).
796: \bibitem{KM80}
797: F.~M.~Kelly and M.~S.~Mathur, Can.~J.~Phys. {\bf 58}, 1416 (1980).
798: \bibitem{Lund}
799: L.~Lundin, B.~Engman, J.~Hilke, and I.~Martinson, Phys.~Scr. {\bf
800: 8}, 274 (1973).
801: \bibitem{PRT}
802: W.~H.~Parkinson, E.~M.~Reeves, and F.~S.~Tomkins, J.~Phys.~B {\bf
803: 9}, 157 (1976).
804: \bibitem{Smith}
805: W.~W.~Smith and A.~Gallagher, Phys.~Rev.~A {\bf 145}, 26 (1966).
806: \bibitem{Hans}
807: W.~J.~Hansen, J.~Phys.~B {\bf 16}, 2309 (1983).
808: \bibitem{God}
809: A.~Godone and C.~Novero, Phys.~Rev.~A {\bf 45}, 1717 (1992).
810: \bibitem{HusR}
811: D.~Husain and G.~J.~Roberts, J.~Chem.~Soc.~Faraday~Trans.~2 {\bf
812: 82}, 1921 (1986).
813: \bibitem{HusS}
814: D.~Husain and J.~Schifino, J.~Chem.~Soc.~Faraday~Trans.~2 {\bf
815: 80}, 321 (1984).
816: \bibitem{Kwong}
817: H.~S.~Kwong, P.~L.~Smith, and W.~H.~Parkinson, Rhys.~Rev.A {\bf
818: 25}, 2629 (1982).
819: \bibitem{Droz}
820: R.~Drozdowski, M.~Ignasiuk, J.~Kwela, and J.~Heldt, Z.~Phys.~D
821: {\bf 41}, 125 (1997).
822: \bibitem{Mitch}
823: C.~Mitchell, J.~Phys.~B {\bf 8}, 25 (1975).
824: \bibitem{Whit}
825: P.~G.~Whitkop and J.~R.~Wiesenfeld, Chem.~Phys.~Lett. {\bf 69},
826: 457 (1980).
827: \bibitem{Kell}
828: J.~F.~Kelly, M.~Harris, and A.~Gallagher, Phys.~Rev.~A {\bf 37},
829: 2354 (1988).
830: \bibitem{Hunter}
831: L.~R.~Hunter, W.~A.~Walker, and D.~S.~Weiss, Phys.~Rev.~Lett. {\bf
832: 56}, 823 (1986).
833: %
834: \bibitem{PorKozRah00} S.G.Porsev, M.G.Kozlov, and Yu.G.Rakhlina,
835: Pis'ma Zh.\ Eksp.\ Theor.\ Fiz.\ {\bf 72}, 862 (2000)[ JETP Lett.,
836: {\bf 72}, 595 (2000)].
837: %
838: \bibitem{DFK}
839: V.~A.~Dzuba, V.~V.~Flambaum, and M.~G.~Kozlov,
840: Pis'ma~Zh.~Eksp.~Teor.~Fiz. {\bf 63}, 844 (1996) [JETP Lett. {\bf
841: 63}, 882 (1996)]; Phys.~Rev.~A {\bf 54}, 3948 (1996).
842: \bibitem{DKPF}
843: V.~A.~Dzuba, M.~G.~Kozlov, S.~G.~Porsev, and V.~V.~Flambaum,
844: Zh.~Eksp.~Theor.~Fiz. {\bf 114}, 1636 (1998) [JETP {\bf 87}, 885
845: (1998)].
846: \bibitem{Din}
847: T.~P.~Dinneen, K.~R.~Vogel, E.~Arimondo, J.~L.Hall, and A.~Gallagher,
848: Phys.~Rev.~A {\bf 59}, 1216 (1999).
849: \bibitem{Kuro}
850: T.~Kurosu and F.~Shimizu, Jpn.~J.~Appl.~Phys., Part 2 {\bf 29},
851: L2127 (1990).
852: \bibitem{Kat}
853: H.~Katori, T.~Ido, Y.~Isoya, and M.~Kuwata-Gonokami,
854: Phys.~Rev.~Lett. {\bf 82}, 1116 (1999).
855: \bibitem{Mg}
856: M.~Machholm, P.~S.~Julienne, and K.-A.~Suominen, Phys.~Rev.~A {\bf
857: 59}, R4113 (1999);
858: \bibitem{DalDav66}
859: A.~Dalgarno and W.~D.~Davidson, Adv.~At.~Mol.~Phys. {\bf 2}, 1
860: (1966).
861: \bibitem{Opt}
862: M.~G.~Kozlov and S.~G.~Porsev, Opt. Spektrosk. {\bf 87}, 384
863: (1999). [Opt. Spectrosc. {\bf 87}, 352 (1999)].
864: \bibitem{Thoul} see, for example,
865: D.~J.~Thouless, {\it The Quantum Mechanics of Many-Body Systems},
866: Chapter IV (Academic, New-York, 1975).
867: \bibitem{PRK1}
868: S.~G.~Porsev, Yu.~G.~Rakhlina, and M.~G.~Kozlov, J.~Phys.~B {\bf
869: 32}, 1113 (1999).
870: \bibitem{PRK2}
871: S.~G.~Porsev, Yu.~G.~Rakhlina, and M.~G.~Kozlov, Phys.~Rev.~A {\bf
872: 60}, 2781 (1999).
873: \bibitem{sidenote} Although the calculations were {\em ab initio} relativistic,
874: for brevity we suppress total angular momentum $j$ in the
875: designations of orbitals.
876: \bibitem{Bogdan}
877: P.~Bogdanovich and G.~\v{Z}ukauskas, Sov.~Phys.~Collection, {\bf
878: 23}, 13 (1983).
879: \bibitem{Moore}
880: C.~E.~Moore, {\it Atomic Energy Levels}, Natl.~Bur.~Stand. (U.S.)
881: Circ. No. 467 (U.S., Washington, 1958).
882: \bibitem{Tiemann} E. Tiemann, private communication.
883: \bibitem{Bausch}
884: C.~W.~Bauschlicher Jr, S.~R.~Langhoff, and H.~Partridge,
885: J.~Phys.~B {\bf 18}, 1523 (1985).
886:
887: \end{thebibliography}
888: %####################################################################
889:
890: \end{document}
891: