physics0404008/qq.tex
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: 
6: \begin{document}
7: 
8: \title{The $\alpha$-dependence of transition frequencies for some ions of 
9: Ti, Mn, Na, C, and O, and the search for variation 
10: of the fine structure constant.}
11: \author{J. C. Berengut}
12: \author{V. A. Dzuba}
13: \email{V.Dzuba@unsw.edu.au}
14: \author{V. V. Flambaum}
15: \email{V.Flambaum@unsw.edu.au}
16: \author{M. V. Marchenko}
17: \affiliation{School of Physics, University of New South Wales, 
18: Sydney 2052,Australia}
19: 
20: \date{\today}
21: 
22: \begin{abstract}
23: 
24: We use the relativistic Hartree-Fock method, many-body perturbation theory
25: and configuration-interaction method to calculate the dependence of 
26:  atomic transition frequencies  on the fine structure constant
27:  $\alpha=e^2/\hbar c$.
28: The results of these calculations will be used in the search 
29: for variation of the fine structure constant in quasar absorption 
30: spectra.
31: 
32: \end{abstract} 
33: 
34: \pacs{PACS: 31.30.Jv, 06.20.Jr 95.30.Dr}
35: \maketitle
36: 
37: The possibility that the fundamental constants vary is suggested by theories
38: unifying gravity with other interactions (see, e.g. \cite{theo1,theo2,theo3}
39: and review \cite{uzan}).
40: The analysis of quasar absorption spectra by means of the many-multiplet
41: method reveals anomalies which can be
42: interpreted in terms of varying fine structure constant $\alpha$
43: \cite{quasar1,quasar2,quasar3}.
44: The first indication that $\alpha$ might have been smaller at early epoch came 
45: from the analysis of magnesium and iron lines \cite{quasar1,quasar2}. 
46: Later inclusion of other lines belonging to many different atoms and ions 
47: (Si, Cr, Ni, Zn, etc.) as well as many samples of data from different
48: gas clouds not only confirmed the initial claim, but made it even stronger
49: \cite{quasar3}. However, there are some recent works in which a similar
50: analysis indicates no variation of $\alpha$ in quasar absorption
51: spectra \cite{Quast,Srianand}. These works use the same many-multiplet
52: method and the results of our calculations of the relativistic effects
53: in atoms, but analyze different samples of data from a different telescope. 
54: It is important to
55: include as much data as possible into the analysis to resolve the
56: differences, and to verify or discard the claim of a varying fine structure
57: constant.
58: 
59: It is natural to analyze fine structure intervals in the search of variation 
60: of $\alpha$. Indeed, initial searches of variation of $\alpha$ in quasar
61: absorption spectra were based on alkali-doublet lines (alkali-doublet method)
62: \cite{AD1,AD2,AD3} and on the fine structure of O~III \cite{Bahcall}.
63: However, all of the present evidence for varying fine structure constant
64: has come from the analysis of the $E1$-transition frequencies (many-multiplet 
65: method) rather than fine structure intervals. 
66: These frequencies are about an order of magnitude more
67: sensitive to the variation of $\alpha$ \cite{quasar2}. However, the corresponding
68: analysis is much more complicated. One needs to perform accurate 
69: {\it ab initio} calculations of the atomic structure to reveal the
70: dependence of transition frequencies on the fine structure constant. We have done such 
71: calculations for many atoms and ions in our previous works 
72: \cite{Dzuba1,Dzuba2}.
73: In the present work we do similar calculations for some other atoms and ions
74: for which data on quasar absorption spectra are available \cite{Murphy},
75: and for which corresponding calculations have not previously been done.
76: 
77: We use the relativistic Hartree-Fock (RHF) method as a starting point of our 
78: calculations. Correlations are included by means of configuration-interaction
79: (CI) method for many valence electron atoms, or by the many-body perturbation
80: theory (MBPT) and Brueckner-orbital method for single valence electron 
81: atoms.
82: The dependence of the frequencies on $\alpha$ is revealed by varying $\alpha$
83: in computer codes.
84: 
85: The results are presented in the form
86: \begin{equation}
87:         \omega=\omega_{0}+qx ,
88: \label{omega}
89: \end{equation}
90: where $ x = (\alpha^{2}/\alpha^{2}_{0})-1$,
91: $\alpha_0$ is the laboratory value of the fine structure 
92: constant, $\omega$ and $\omega_{0}$ are the frequencies of the transition in
93: quasar absorption spectra and in the laboratory, respectively, and $q$ is 
94: the relativistic energy shift that comes from the calculations.
95: Comparing the laboratory frequencies, $\omega_0$, with those measured in
96: the quasar absorption spectra, $\omega$, allows one to obtain the value
97: of $\alpha$ billions of years ago.
98: 
99: The method of calculations is described in detail in our early works
100: \cite{Dzuba1,Dzuba2}. Here we only discuss the details specific for current
101: calculations.
102: 
103: Some atoms and ions considered in the present work represent open-shell (many valence electron)
104: systems. Therefore, the Hartree-Fock procedure needs to be further specified.
105: The natural choice is to remove all open-shell electrons and start the
106: Hartree-Fock calculations for the closed-shell core. 
107: However, this usually leads to poor convergence of the subsequent CI method. 
108: Better convergence can be achieved using the so called $V^{N-1}$ approximation
109: in which only one valence electron is removed.
110: Since we calculate not only the ground state but also excited states of different
111: configurations, it is convenient to remove the electron
112: which changes its state in the transition. Single-electron basis states
113: for valence electrons are calculated in the $V^{N-1}$ potential of the 
114: frozen-core.
115: 
116: The $V^{N-1}$ potential corresponds to an open-shell system.
117: We include the contribution of the open shells
118: into the Hartree-Fock potential as if they were totally filled and then
119: multiply them by a weighting coefficient. Note that this procedure must
120: not destroy the cancellation of the self-action (we would like to remind the
121: reader that there is exact cancellation between direct and exchange 
122: self-action in the Hartree-Fock equations for the closed-shell systems).
123: 
124: For the CI calculations we use B-splined single-electron basis set similar 
125: to those developed by Johnson {\it et al} \cite{BS1,BS2,BS3}.
126: The main difference is that we use the open-shell RHF Hamiltonian described 
127: above to calculate the B-splined states.
128: 
129: There are two major sources of inaccuracy in the standard CI calculations. 
130: One is incompleteness of the basis set and another is core-valence 
131: correlations.
132: We use a fitting procedure to model both effects.
133: We add an extra term into a single-electron part of the Hamiltonian for
134: the valence electrons:
135: \begin{equation}
136:   U(r)=-\frac{\alpha_c}{2\left(r^4+a^4 \right)}.
137: \label{fit}
138: \end{equation}
139: Here $\alpha_c$ is the polarizability of the atomic core and $a$ is a cut-off
140: parameter that is introduced to remove the singularity at $r=0$. 
141: We use $a=a_b$ (Bohr radius) and treat $\alpha_c$ as a fitting parameter.
142: The values of $\alpha_c$ for each partial wave ($s,p,d$) are chosen to fit
143: the experimental energy levels of the many-electron atom.
144: 
145: The term (\ref{fit}) describes polarization of the atomic core by valence 
146: electrons. It can be considered as a semi-empirical approximation to
147: the correlation interaction of a particular valence electron with the core.
148: It also allows us to improve the convergence of the CI calculations by
149: modifying the single-electron basis states. Our calculations for rare-earth
150: ions \cite{Safronova1,Safronova2} have demonstrated that using this term 
151: allows one to obtain 
152: good accuracy of calculations with the minimum number of single-electron
153: basis states (one in each partial wave in the cited works).
154: 
155: Below we present the details and results of calculations for the atoms and ions 
156: considered.
157: All transition frequencies are presented with respect to the ground state.
158: Therefore we use the term ``energy levels'' instead. If a transition
159: between excited states is needed, the corresponding relativistic energy
160: shift $q$ is the difference between the level shifts
161: ($q_{2 \rightarrow 1} = q_2 - q_1$).
162: 
163: \paragraph{Manganese ($Z=25$):}  The ground state of Mn$^+$ is 
164: $3d^{5}4s \ ^7S_{3}$ and we need to consider transitions into the $3d^{4}4s4p$
165: configuration. Earlier we also considered transitions to the states of the
166: $3d^54p$ configuration \cite{Dzuba1}. Since in the present work we use 
167: different basis set, we have repeated calculations for this configuration in order
168: to check their accuracy.
169: 
170: The RHF calculations are done in the $V^{N-1}$ approximation with the
171: $3d^5$ configuration of external electrons. The $4s, 4p$ and higher states are
172: calculated in the same $V^{N-1}$ potential. We use $\alpha_c = 2.05 a_B^3$
173: for the $p$-wave as a fitting parameter (see formula (\ref{fit})).
174: The results are presented in Table \ref{Mn}.
175: Fitting changes both energies and $q$-coefficients by less than 10\%,
176: and agreement with previous calculations is also within 10\%.
177: Therefore, we use 10\% as a conservative estimate of the accuracy of $q$. 
178: 
179: Note that the relativistic
180: shift is positive for the $s-p$ singe-electron transitions and negative
181: for the $d-p$ transitions. Having transitions with different signs of $q$-coefficients
182: in the same atom (ion) helps to fight
183: systematic errors in the search for variation of $\alpha$ 
184: (see Ref.~\cite{Dzuba1} for details). 
185: 
186: 
187: \begin{table}
188: \caption{Energies and relativistic energy shifts ($q$) for Mn$^+$ (cm$^{-1}$)}
189: \label{Mn}
190: \begin{ruledtabular}
191: \begin{tabular}{llcccrr}
192: \multicolumn{2}{c}{State} & \multicolumn{3}{c}{Energy}& 
193: \multicolumn{2}{c}{$q$} \\
194: &&\multicolumn{2}{c}{theory} & experiment & \\
195: &&no fitting & fitted & \cite{Moore} & this work & \cite{Dzuba2} \\ 
196: \hline
197:  $3d^5 4p$   & $^7P_2$    & 36091 & 38424 & 38366 &   869 &  918 \\
198:  $3d^5 4p$   & $^7P_3$    & 36252 & 38585 & 38543 &  1030 & 1110 \\
199:  $3d^5 4p$   & $^7P_4$    & 36483 & 38814 & 38807 &  1276 & 1366 \\
200:  $3d^4 4s4p$ & $^7P_2$    & 97323 & 83363 & 83255 & -3033 & \\
201:  $3d^4 4s4p$ & $^7P_3$    & 97554 & 83559 & 83376 & -2825 & \\
202:  $3d^4 4s4p$ & $^7P_4$    & 97858 & 83818 & 83529 & -2556 & \\
203: \end{tabular}
204: \end{ruledtabular}
205: \end{table}
206: 
207: \paragraph{Titanium( $Z=22$):} We perform calculations for both Ti$^+$ 
208: and Ti$^{2+}$ starting from the same RHF approximation, and 
209: using the same single-electron basis set.
210: The ground state of Ti$^+$ is $3d^{2}4s \ ^4F_{3/2}$
211: and we need to consider transitions into states of the $3d^{2}4p$
212: configuration.
213: The ground state of Ti$^{2+}$ is $3d^{2} \ ^3F_{2}$
214: and we need to consider transitions into the states of the $3d4p$
215: configuration.
216: Therefore it is convenient to do the
217: RHF calculations for the Ti$^{2+}$ ion
218: with the $3d^{2}$ open-shell configuration. The $4s$, $4p$ and other
219: basis states for the CI method are calculated in the frozen-core field 
220: of Ti$^{2+}$. 
221: 
222: The fitting parameters chosen are $\alpha_c=0.38 a_B^3$ for 
223: $s$-electrons and $\alpha_c=0.065 a_B^3$ for $d$-electrons.
224: The results are presented in Table \ref{Ti}.
225: As in the case of Mn$^+$, there are negative and positive relativistic shifts.
226: The effects of fitting and change of basis set does not exceed 10\%.
227: The values of the $q$-coefficients for titanium are consistent with 
228: calculations for other atoms and with semi-empirical estimations
229: using the formulas presented in \cite{Dzuba1}. 
230: In particular, the values of the negative $q$-coefficients for 
231: the $d-p$ transitions are very close to the values for similar transitions in 
232: Cr~II \cite{Dzuba1}. The positive coefficients for Ti$^+$ are very close 
233: to those for Mn$^+$ after rescaling by $Z^2$ according to the semi-empirical
234: formula \cite{Dzuba1}.
235: 
236: \begin{table}
237: \caption{Energies and relativistic energy shifts ($q$) for Ti$^+$ and Ti$^{2+}$ 
238: (cm$^{-1}$)}
239: \label{Ti}
240: \begin{ruledtabular}
241: \begin{tabular}{llcccr}
242: \multicolumn{2}{c}{State} &   \multicolumn{3}{c}{Energy}& $q$ \\
243: &&\multicolumn{2}{c}{theory} & experiment & \\
244: &&no fitting & fitted & \cite{Moore} & \\ 
245: \hline
246: \multicolumn{6}{c}{Ti$^+$} \\
247:  $3d^2 4p$ & $^4G_{5/2}$  & 27870 & 29759 & 29544 &   396 \\
248:  $3d^2 4p$ & $^4F_{3/2}$  & 28845 & 30691 & 30837 &   541 \\
249:  $3d^2 4p$ & $^4F_{5/2}$  & 28965 & 30813 & 30959 &   673  \\
250:  $3d^2 4p$ & $^4D_{1/2}$  & 30582 & 32416 & 32532 &   677  \\
251:  $3d^2 4p$ & $^4D_{3/2}$  & 30670 & 32510 & 32603 &   791  \\
252:  $3d4s4p$  & $^4D_{1/2}$  & 50651 & 52185 & 52330 & -1564  \\
253: \multicolumn{6}{c}{Ti$^{2+}$} \\
254:  $3d4p$    & $^3D_1$      & 80558 &       & 77000 & -1644  \\
255: \end{tabular}
256: \end{ruledtabular}
257: \end{table}
258: 
259: \paragraph{Sodium ($Z=11$):} In contrast to the ions considered above,
260: sodium is an atom with one external electron above closed
261: shells. Its ground state is $1s^2 2s^2 2p^6 3s \ ^2S_{1/2}$.
262: Very accurate calculations are possible for such systems by
263: including certain types of correlation diagrams to all orders (see, e.g.
264: \cite{Dzuba89,Johnson91}). However, since both relativistic and correlation
265: effects for sodium are small we use a simplified approach. We calculate
266: the correlation potential $\hat \Sigma$ (the average value of this operator
267: is the correlation correction to the energy of the external electron) 
268: in the second order only.
269: Then we use it to modify the RHF equations for the valence electron and
270: to calculate the so called Brueckner-orbitals. Note that due to iterations
271: of $\hat \Sigma$ certain types of correlation diagrams are still included
272: in all orders in this procedure. The final accuracy of the energy is better
273: than 1\%, and for the fine structure accuracy is 2-6\% (see Table \ref{Na}).
274: We believe that the accuracy for the relativistic shifts $q$ is on the 
275: same level.
276: 
277: \begin{table}
278: \caption{Energies and relativistic energy shifts ($q$) for Na (cm$^{-1}$)}
279: \label{Na}
280: \begin{ruledtabular}
281: \begin{tabular}{llccr}
282: \multicolumn{2}{c}{State} &   \multicolumn{2}{c}{Energy}& $q$ \\
283: &&\multicolumn{1}{c}{theory} & experiment \cite{Moore} & \\
284: \hline
285:  $3p$      & $^2P_{1/2}$  & 16858        & 16956 &    45  \\
286:  $3p$      & $^2P_{3/2}$  & 16876        & 16973 &    63  \\
287:  $4p$      & $^2P_{1/2}$  & 30124        & 30267 &    53  \\
288:  $4p$      & $^2P_{3/2}$  & 30130        & 30273 &    59  \\
289: \end{tabular}
290: \end{ruledtabular}
291: \end{table}
292: 
293: \begin{table}
294: \caption{Energies and relativistic energy shifts ($q$) for the carbon atom and its ions (cm$^{-1}$)}
295: \label{carbon}
296: \begin{ruledtabular}
297: \begin{tabular}{llrrr}
298: \multicolumn{2}{c}{State} &   \multicolumn{2}{c}{Energy}& $q$ \\
299: &&\multicolumn{1}{c}{theory} & \multicolumn{1}{c}{experiment \cite{Moore}} & \\
300: \hline
301: \multicolumn{5}{c}{C} \\
302: $2s 2p^3$    & $^3D_3$   & 66722   & 64087   &      151     \\
303: $2s 2p^3$    & $^3D_1$   & 66712   & 64090   &      141     \\
304: $2s 2p^3$    & $^3D_2$   & 66716   & 64091   &      145     \\     
305: $2s 2p^3$    & $^3P_1$   & 75978   & 75254   &      111     \\
306: $2s 2p^3$    & $^3S_1$   &100170   &105799   &      130     \\
307: \multicolumn{5}{c}{C$^+$} \\
308: $2s^2 2p$   & $^2P_{1/2}$ &    74  &    63   &       63      \\
309: $2s 2p^2$   & $^2D_{5/2}$ & 76506  & 74930   &      179      \\
310: $2s 2p^2$   & $^2D_{3/2}$ & 76503  & 74933   &      176      \\
311: $2s 2p^2$   & $^2S_{1/2}$ & 97993  & 96494   &      161      \\
312: \multicolumn{5}{c}{C$^{2+}$} \\
313: $2s 2p$   & $^1P_1$   &    104423  & 102352  &    162      \\
314: \multicolumn{5}{c}{C$^{3+}$} \\
315: $2p$   & $^2P_{1/2}$   &    65200  &  64484  &    104      \\
316: $2p$   & $^2P_{3/2}$   &    65328  &  64592  &    232      \\
317: \end{tabular}
318: \end{ruledtabular}
319: \end{table}
320: 
321: \paragraph{Carbon ($Z=6$):} Relativistic effects for carbon and its ions 
322: are small and calculations can be done without fitting parameters.
323: The ground state of neutral carbon is $1s^2 2s^2 2p^2 \ ^3P_0$.
324: Our RHF calculations for this atom include all electrons,
325: however, since we need to consider configurations with excitations 
326: from both $2s$ and $2p$ states, we treat both as valence states in CI.
327: 
328: For neutral carbon we have performed the calculations for the ground
329: state configuration as well as for excited configurations
330: $2s^2 2p3s$, $2s 2p^3$, $2s^2 2p4s$,$2s^2 2p3d$, $2s^2 2p4d$,
331: $2s^2 2p5d$ and $2s^2 2p6d$. However, we present in Table~\ref{carbon}
332: only results for the $2s 2p^3$ configuration. The relativistic energy
333: shift for all other configurations is small ($q < 50\ \rm{cm}^{-1}$).
334: This is smaller than uncertainty of the $q$-coefficients for heavier atoms
335: and ions. Since the analysis of quasar spectra is based on comparison of the 
336: relativistic effects in light and heavy atoms (ions), small relativistic energy shifts
337: in light atoms can be neglected. The $q$-coefficients for the $2s 2p^3$
338: configuration are larger because this configuration corresponds to the
339: $2s - 2p$ transition from the ground state. These are the lowest valence 
340: single-electron states with the largest relativistic effects. Other
341: excited configurations correspond to the $2p - ns$ or $2p - nd$ ($n \geq 3$)
342: transitions. However, relativistic energy shifts for higher states are smaller
343: \cite{Dzuba1}.
344: 
345: The calculations for C$^{2+}$ and C$^{3+}$ are done in the potential of the 
346: closed-shell (helium) core. As can be seen from Table \ref{carbon}, accuracy 
347: for the energies is within 10\%. We estimate the accuracy of $q$-coefficients 
348: at around 10-20\%.
349: 
350: \begin{table}
351: \caption{Energies and relativistic energy shifts ($q$) for oxygen ions (cm$^{-1}$)}
352: \label{O}
353: \begin{ruledtabular}
354: \begin{tabular}{llrrr}
355: \multicolumn{2}{c}{State} &   \multicolumn{2}{c}{Energy}& $q$ \\
356: &&\multicolumn{1}{c}{theory} & \multicolumn{1}{c}{experiment \cite{Moore}} & \\
357: \hline
358: \multicolumn{5}{c}{O$^+$} \\
359: $2s 2p^4$    & $^4P_{5/2}$   & 122620   & 119873   &      346     \\
360: $2s 2p^4$    & $^4P_{3/2}$   & 122763   & 120000   &      489     \\
361: $2s 2p^4$    & $^4P_{1/2}$   & 122848   & 120083   &      574     \\
362: \multicolumn{5}{c}{O$^{2+}$} \\
363: $2s 2p^3$    & $^3D_{1}$    &    121299 & 120058   &      723      \\
364: $2s 2p^3$    & $^3P_{1}$    & 143483    & 142382   &      726      \\
365: \multicolumn{5}{c}{O$^{3+}$} \\
366: $2s 2p^2$   & $^2D_{3/2}$   &   129206  & 126950  &       840      \\
367: \multicolumn{5}{c}{O$^{5+}$} \\
368: $1s^2 2p$   & $^2P_{1/2}$   &    97313  &  96375  &       340      \\
369: $1s^2 2p$   & $^2P_{3/2}$   &    97913  &  96908  &       872      \\
370: \end{tabular}
371: \end{ruledtabular}
372: \end{table}
373: 
374: \paragraph{Oxygen ($Z=8$):}
375: 
376: Relativistic effects for oxygen ions are comparatively large, 
377: and become larger with increasing electric charge. 
378: This is in agreement with semi-empirical formulae presented in \cite{Dzuba1}.
379: For neutral oxygen, however, $q$-coefficients are approximately 20 cm$^{-1}$ or less; 
380: these results are not presented here.
381: 
382: \vspace{10pt}
383: This work was supported in part by the Australian Research Council.
384: 
385: \bibliography{qq}
386: 
387: \end{document}
388: 
389: