1: \documentclass[aps,twocolumn,pra,showpacs,floatfix]{revtex4}
2: %\documentclass[preprint,pra,floatfix]{revtex4}
3:
4: %\usepackage[dvips]{graphicx}
5: \usepackage{epsfig}
6: \usepackage{graphicx}
7: \usepackage{dcolumn}
8: \usepackage{amsmath}
9:
10: \begin{document}
11:
12: %\preprint{}
13:
14: \title{Relativistic corrections to transition frequencies of Fe~I and
15: search for variation of the fine structure constant}
16:
17: \author{V. A. Dzuba}
18: \affiliation{School of Physics, University of New South Wales,
19: Sydney 2052, Australia}
20: \author{V. V. Flambaum}
21: \affiliation{School of Physics, University of New South Wales,
22: Sydney 2052, Australia}
23:
24: \date{\today}
25:
26: \begin{abstract}
27:
28: Relativistic energy shifts of the low energy levels of Fe
29: have been calculated using the Dirac-Hartree-Fock and
30: configuration interaction techniques. The results are to be used in
31: the search for the space-time variation of the fine structure constant
32: in quasar absorption spectra. The values of the shifts are the largest
33: among those used in the analysis so far. This makes Fe a good candidate
34: for the inclusion into the analysis.
35:
36: \end{abstract}
37:
38: \pacs{PACS: 31.30.Jv, 06.20.Kr, 95.30.Dr}
39:
40: \maketitle
41:
42: \section{Introduction}
43:
44: Theories unifying gravity with other interactions suggest a possibility
45: of temporal and spatial variations of the fundamental constants of nature;
46: reviews of these theories and results of measurement can be found
47: in Refs.~\cite{Uzan,Flambaum07a}. Strong evidence that the fine-structure
48: constant might be smaller about ten billion years ago was found in the
49: analysis of quasar absorption
50: spectra~\cite{Webb99,Webb01,Murphy01a,Murphy01b,Murphy01c,Murphy01d}.
51: This result was obtained using the data from the Keck
52: telescope in Hawaii. However, an analysis of the data
53: from the VLT telescope in Chile, performed by different groups~\cite{vlt1,vlt2}
54: gave a null result. There is an outgoing debate
55: in the literature about possible reasons for the disagreement
56: (see. e.g.~\cite{Flambaum07,Srianand07}).
57:
58: All these results were obtained with the use of the so called
59: many-multiplet method suggested in Ref.~\cite{Dzuba99}.
60: This method requires calculation of relativistic corrections
61: to frequencies of atomic transitions to reveal their dependence
62: on the fine-structure constant. Atomic calculations for a large
63: number of transitions in many atoms and ions of astrophysical
64: interest were reported in Refs.~\cite{Dzuba99a,Dzuba01,Dzuba02,
65: Berengut04,Berengut05,Berengut06,archDzuba,Porsev,Johnson,Savukov}.
66: In present paper we perform similar calculations for neutral iron
67: which was never considered before. Until very recent we were unaware
68: about any lines of neutral iron observed in the quasar absorption
69: spectra. It was Prof. P. Molaro~\cite{Molaro} who brought to our
70: attention the existence of such data and the intention of his group
71: to use them in the analysis.
72:
73: Using Fe~I in the search for variation of the fine structure constant
74: has several advantages. First, the values of the relativistic energy
75: shifts are large due to relatively larger nuclear charge ($Z$=26)
76: and strong configuration mixing. Second, these values vary strongly
77: from state to state which makes it hard to mimic the effect of varying
78: fine structure constant by any systematics. Finally, iron is very
79: abundant element in the universe. Great part of the previous analysis
80: was performed using the data from Fe~II.
81:
82: Calculations for Fe~I are difficult due to large number of valence electrons
83: and strong configuration mixing. Its ground state configuration for
84: outermost electrons is $3d^64s^2$ which is strongly mixed with the
85: $3d^74s$ configuration. There is strong configuration mixing for the
86: $3d^64s4p$ and $3d^74p$ odd-parity configurations for the excited
87: states. With eight valence electrons and strong configuration mixing
88: full-scale accurate {\em ab initio} calculations for Fe~I would require
89: enormous computer power. We have chosen a different approach. Below
90: we report a simple method which is specially designed for systems
91: with strong mixing of several distinct configurations. It combines
92: {\em ab initio} Hartree-Fock and configuration interaction (CI)
93: techniques with some semi-empirical fitting and gives very
94: reasonable results at very low cost in terms of computer power.
95: The approach is similar to the well-known multi-configuration
96: relativistic Hartree-Fock method (see, e.g.~\cite{Grant}) and
97: can probably be considered as a simple version of it.
98: The accuracy for the energy levels of Fe~I is within few per cent
99: of experimental values while estimated accuracy for the relativistic
100: energy shifts is on the level of 20 to 30\%.
101: Due to strong configuration mixing the results are sensitive to the
102: distances between energy levels. Therefore, special care has been
103: taken to reproduce experimental positions of the energy levels.
104:
105: \section{Method}
106:
107: It is convenient to present the dependence of atomic frequencies on
108: the fine-structure constant $\alpha$ in the vicinity of its physical
109: value $\alpha_0$ in the form
110: \begin{equation}
111: \omega(x) = \omega_0 + qx,
112: \label{omega}
113: \end{equation}
114: where $\omega_0$ is the laboratory value of the frequency and
115: $x = (\alpha/\alpha_0)^2-1$, $q$ is the coefficient which is to be
116: found from atomic calculations. Note that
117: \begin{equation}
118: q = \left .\frac{d\omega}{dx}\right|_{x=0}.
119: \label{qq}
120: \end{equation}
121: To calculate this derivative numerically we use
122: \begin{equation}
123: q \approx \frac{\omega(+\delta) - \omega(-\delta)}{2\delta}.
124: \label{deriv}
125: \end{equation}
126: Here $\delta$ must be small to exclude non-linear in $\alpha^2$ terms.
127: In the present calculations we use $\delta = 0.05$, which leads to
128: \begin{equation}
129: q \approx 10 \left(\omega(+0.05) - \omega(-0.05)\right).
130: \label{deriv05}
131: \end{equation}
132: To calculate the coefficients $q$ using (\ref{deriv05}), $\alpha$ must be varied
133: in the computer code. Therefore, it is convenient to use a form
134: of the single electron wave function in which the dependence on $\alpha$ is
135: explicitly shown (we use atomic units in which $e=\hbar=1, \alpha = 1/c$)
136: \begin{equation}
137: \psi(r)_{njlm}=\frac{1}{r}\left(\begin {array}{c}
138: f_{v}(r)\Omega(\mathbf{n})_{\mathit{jlm}} \\[0.2ex]
139: i\alpha g_{v}(r) \widetilde{ \Omega}(\mathbf{n})_{\mathit{jlm}}
140: \end{array} \right),
141: \label{psi}
142: \end{equation}
143: where $n$ is the principal quantum number and an index $v$
144: replaces the three-number set $n,j,l$.
145: This leads to a form of radial equation for single-electron
146: orbitals which also explicitly depends on $\alpha$:
147: \begin{equation}
148: \begin {array}{c} \dfrac{df_v}{dr}+\dfrac{\kappa_{v}}{r}f_v(r)-
149: \left[2+\alpha^{2}(\epsilon_{v}-\hat{V}_{HF})\right]g_v(r)=0, \\[0.5ex]
150: \dfrac{dg_v}{dr}-\dfrac{\kappa_{v}}{r}f_v(r)+(\epsilon_{v}-
151: \hat{V}_{HF})f_v(r)=0, \end{array}
152: \label{Dirac}
153: \end{equation}
154: here $\kappa=(-1)^{l+j+1/2}(j+1/2)$,
155: and $\hat{V}_{HF}$ is the Hartree-Fock potential.
156: Equation (\ref{Dirac}) with $\alpha = \alpha_0 \sqrt{\delta +1}$
157: and different Hartree-Fock potential $\hat{V}_{HF}$ for
158: different configurations is used to construct single-electron orbitals.
159:
160: \begin{table}
161: \caption{Even and odd configurations of Fe and effective core
162: polarizability $\alpha_p$ (a.u.) used in the calculations.}
163: \label{sets}
164: \begin{ruledtabular}
165: \begin{tabular}{l l l l}
166: \multicolumn{1}{c}{Set} &\multicolumn{1}{c}{Parity} &
167: \multicolumn{1}{c}{Configuration} &
168: \multicolumn{1}{c}{$\alpha_p$} \\
169: \hline
170: 1 & Even & $3d^64s^2$ & 0.4 \\
171: 2 & Even & $3d^74s$ & 0.4192 \\
172: 3 & Even & $3d^64p^2$ & 0.4 \\
173: 4 & Even & $3d^8$ & 0.465 \\
174: \hline
175: 5 & Odd & $3d^64s4p$ & 0.39 \\
176: 6 & Odd & $3d^74p$ & 0.412 \\
177: 7 & Odd & $3d^54s^24p$ & 0.409 \\
178: \end{tabular}
179: \end{ruledtabular}
180: \end{table}
181:
182: Table~\ref{sets} lists configurations considered in present work. First four
183: are even configurations and other three are odd configurations.
184: %We have
185: %checked that other configurations (e.g., $3d^8$) give only negligible
186: %contribution to the states of interest and we do not include them.
187: We perform self-consistent Hartree-Fock calculations for each configuration
188: separately. This allows to account for the fact that single-electron
189: states depend on the configurations. For example, the $3d$ state in the
190: $3d^64s^2$ configuration is not the same as the $3d$ state in the $3d^74s$
191: configuration. In principle, it is possible to account for these
192: differences in the CI calculations. One would need to have a complete set
193: of single-electron states and construct many-electron basis states by
194: redistributing valence electrons over these single-electron basis states.
195: Then actual many-electron states are found by diagonalization of matrix
196: of the effective CI Hamiltonian. This approach works very well for the case
197: of two or three valence electrons (see, e.g.~\cite{JETP,Kozlov96,Johnson98}).
198: However, for eight valence electrons it would lead to a matrix of enormous
199: size making it practically impossible to saturate the basis while using limited
200: computer power. The results with unsaturated basis are very unstable and
201: strongly depend on where the basis is truncated. Therefore, we prefer
202: to account for the differences in the configurations on the Hartree-Fock
203: rather than CI stage of the calculations.
204:
205: The self-consistent Hartree-Fock procedure is done for every configuration
206: listed in Table~\ref{sets} separately. Then valence states found in the
207: Hartree-Fock calculations are used as basis states for the CI calculations.
208: It is important for the CI method that atomic core ($1s^2 \dots 3p^6$) remains
209: the same for all configurations. We use the core which corresponds to the
210: ground state configuration. Change in the core due to change of the valence
211: state is small and can be neglected. This is because core states are not
212: sensitive to the potential from the electrons which are on large distances
213: (like $4s$ and $4p$ electrons). The $3d$ electrons are on smaller distances
214: and have larger effect on atomic core. However, in most of the cases
215: (see Table~\ref{sets}) only one among six $3d$ electrons change its state.
216: Therefore their effect on atomic core is also small. More detailed
217: discussion on the effect of valence electrons on atomic core can be
218: found in Refs.~\cite{VN,VN1}.
219:
220: All configurations in Table~\ref{sets} correspond to an open-shell
221: system. We perform the calculations staying within central-field
222: approximation but using fractional occupation numbers. As a result
223: we have 23 singe-electron basis states for valence electrons:
224: $3d^{(i)}_{3/2},3d^{(i)}_{5/2},4s^{(i)},4p^{(i)}_{1/2},4p^{(i)}_{3/2}$.
225: Here index $i$ is the set number (as in Table~\ref{sets}).
226: Note that total number of basis states is less than 5 times number
227: of sets
228: because many configurations don't include particular single-electron
229: states. Note also that our basis set is non-orthogonal, e.g.
230: $0 < \langle 3d^{(i)}_{3/2}|3d^{(j)}_{3/2} \rangle <1$. The implications
231: of this fact will be discussed below.
232:
233: The effective Hamiltonian for valence electrons has the form
234: \begin{equation}
235: \hat H^{\rm eff} = \sum_{i=1}^8 \hat h_{1i} +
236: \sum_{i < j}^8 e^2/r_{ij},
237: \label{heff}
238: \end{equation}
239: $\hat h_1(r_i)$ is the one-electron part of the Hamiltonian
240: \begin{equation}
241: \hat h_1 = c \mathbf{\alpha \cdot p} + (\beta -1)mc^2 - \frac{Ze^2}{r}
242: + V_{core} + \delta V.
243: \label{h1}
244: \end{equation}
245: Here $\mathbf{\alpha}$ and $\beta$ are Dirac matrixes, $V_{core}$ is
246: Hartree-Fock potential due to 18 core electrons ($1s^2 \dots 3p^6$)
247: and $\delta V$
248: is the term which simulates the effect of the correlations between core
249: and valence electrons. It is often called {\em polarization potential} and
250: has the form
251: \begin{equation}
252: \delta V = - \frac{\alpha_p}{2(r^4+a^4)}.
253: \label{dV}
254: \end{equation}
255: Here $\alpha_p$ is polarization of the core and $a$ is a cut-off parameter
256: (we use $a = a_B$).
257: The form of the $\delta V$ is chosen to coincide with the standard polarization
258: potential on large distances ($-\alpha_p/2r^4$). However we use it on distances
259: where valence electrons are localized. This distances are not large, especially
260: for the $3d$ electrons. Therefore we consider $\delta V$ as only rough
261: approximation to real correlation interaction between core and valence
262: electrons and treat $\alpha_p$ as fitting parameters. The values of $\alpha_p$
263: for each configuration of interest are presented in Table~\ref{sets}.
264: They are chosen to fit the experimental position of the configurations
265: relative to each other. The value of $\alpha_p$ for the $3d^64p^2$
266: configuration is taken to be the same as for the ground state configuration
267: because actual position of this configuration in the
268: energy spectrum is not known. For all configurations the values of
269: $\alpha_p$ are very close. This is not a surprise since the core is
270: always the same. One can probably say that small difference in $\alpha_p$
271: for different configurations simulates the effect of incompleteness of the
272: basis and other imperfections in the calculations.
273:
274: \subsection{CI calculations with a non-orthogonal basis.}
275:
276: As it was mentioned above we have a set of single-electron states which is
277: not orthogonal. The $3d$, $4s$ and $4p$ states in the configurations listed
278: in Table~\ref{sets} are similar but not the same. In principle, it may lead
279: to complication in the CI procedure, starting from non-orthogonality
280: of many-electron basis states which would lead in turn to complications in
281: calculation of matrix elements and matrix diagonalization.
282: However, most of these complications can be avoided by appropriate selection of
283: the configurations included in the calculations. It is sufficient to
284: obey the two rules:
285: \begin{itemize}
286: \item Forbid configurations which have singe-electron states taken from different
287: sets, e.g. $3d^m_i3d^n_j4s_k4s_l$. Here $i,j,k$ and $l$ are set numbers as
288: in Table~\ref{sets} ($i \neq j$ or/and $k \neq l$)
289: and $m$ and $n$ are number of electrons in each of the
290: $3d$ state ($m+n=6$).
291: \item Don't generate additional configurations by exciting electrons
292: to the orbitals of the same symmetry, e.g. $3d^64s^2 \rightarrow 3d^64s5s$.
293: \end{itemize}
294: In present calculations we use only those configurations which are listed
295: in Table~\ref{sets}.
296:
297: If all single-electron states for every many-electron
298: basis state are taken from the same set then the many-electron basis states
299: remain orthogonal to each other. Indeed, states of the same configuration
300: are orthogonal to each other as in the standard CI technique. States of
301: different configurations are orthogonal because at least one electron
302: changes its angular symmetry in the transition between the configurations.
303: For example all states of the $3d^64s^2$ configuration are orthogonal to
304: all states of the $3d^74s$ configuration because of the $s - d$ transition
305: involved.
306:
307: Since many-electron basis functions remain orthogonal matrix diagonalization
308: is not affected. Calculation of the matrix elements between states of the same
309: configuration is not affected as well. The only part of the CI procedure
310: which is affected is calculation of matrix elements between basis states
311: of different configurations. Here single electron part $\hat h_1$ (\ref{h1})
312: of the Hamiltonian does not contribute because this is a scalar operator
313: which cannot change angular symmetry of a single-electron state.
314: Only Coulomb integrals contribute to the matrix elements and
315: their calculation must be accomplished by the product
316: of overlaps between similar states from different sets. For example,
317: Coulomb interaction between the $3d^64s^2$ and $3d^64p^2$ configurations
318: has the form (in non-relativistic notations):
319: \[
320: F_1(4s_1,4p_3,4s_1,4p_3)\langle 3d_1|3d_3 \rangle^6.
321: \]
322: Here $F_1$ is dipole Coulomb integral, indexes 1 and 3 numerate basis sets
323: as in Table~\ref{sets}, $\langle 3d_1|3d_3 \rangle$ is the overlap between
324: different $3d$ functions.
325:
326: \section{Results and discussion}
327:
328: Neutral iron is an interesting system as a challenge for the calculations
329: and as a candidate for the search of the variation of the fine structure
330: constant. There is strong configuration mixing between the $3d^64s^2$
331: and the $3d^74s$ even configurations in the ground state and the
332: $3d^64s4p$ and the $3d^74p$ configurations for the odd excited states.
333: The latter mixing is a fortunate feature which makes Fe~I a convenient
334: object for the analysis. Let us elaborate. It is important to have
335: relativistic frequency shifts of the atomic transitions used in the analysis
336: to be as large as possible. The value of the shift depends on how many
337: electrons change their states in the transitions and how large is the
338: change of electron momentum in each single-electron transition
339: (see, e.g.~\cite{Dzuba99a}). The transition between $3d^64s^2$ and
340: $3d^64s4p$ configurations is basically a $4s - 4p$ transition.
341: However, mixing with the $3d^74s$ configuration in the upper state
342: adds one more single-electron transition ($4s - 3d$) and makes the
343: frequency shift larger. Note that the presence of both the $3d^64s4p$
344: and the $3d^74p$ configurations is important. The first configuration
345: is needed for the
346: transition to the ground state to be strong electric dipole transition,
347: otherwise it will not be observed. The second configuration is needed for
348: the relativistic frequency shift to be large. It is fortunate that
349: strong configuration mixing between these two configurations takes
350: place for most of the low odd states of Fe~I.
351:
352: On the other hand this strong configuration mixing is a big challenge
353: for the calculations. It makes the results for the relativistic
354: energy shifts (the $q$-coefficients) to be unstable since they are
355: very sensitive to the value of the mixing. Note that the configuration
356: mixing in the ground state is also important. The admixture of the
357: $3d^74s$ configuration adds the contribution of the $3d - 4p$
358: transition to the relativistic frequency shift. This contribution
359: has an opposite sing as compared to the $4s - 4p$ transition
360: which adds to the instability of the results.
361:
362: Since configuration mixing is very sensitive to the energy intervals
363: between the states the most reliable results can be obtained in the
364: calculations which reproduce correctly experimental spectrum.
365: In present calculations this is achieved with the use of the
366: core polarization term (\ref{dV}) in the Hamiltonian and fitting
367: the data by changing the core polarizability parameter $\alpha_p$.
368: Note however that only fine tuning was needed since in the end
369: the values of the $\alpha_p$ for different configurations turned to
370: be very close to each other (see Table~\ref{sets}).
371:
372: \begin{figure}
373: \centering
374: \epsfig{figure=f3.eps,scale=0.4}
375: \caption{Odd-parity energy levels of Fe~I with total momentum $J=3$ as
376: functions of the fine structure constant.}
377: \label{f3}
378: \end{figure}
379:
380:
381: \begin{figure}
382: \centering
383: \epsfig{figure=f4.eps,scale=0.4}
384: \caption{Odd-parity energy levels of Fe~I with total momentum $J=4$ as
385: functions of the fine structure constant.}
386: \label{f4}
387: \end{figure}
388:
389:
390: \begin{figure}
391: \centering
392: \epsfig{figure=f5.eps,scale=0.4}
393: \caption{Odd-parity energy levels of Fe~I with total momentum $J=5$ as
394: functions of the fine structure constant.}
395: \label{f5}
396: \end{figure}
397:
398: Another source of possible numerical instability of the results
399: for particular states is level pseudo-crossing
400: (see, e.g.~\cite{Dzuba01,Dzuba02}). Energies of the states when
401: considered as function of $\alpha^2$ may come close to each other
402: in the vicinity of the physical value of $\alpha$. Then small
403: error in the position of level crossing may lead to large
404: error in the $q$-coefficient which is actually the slop
405: of the curve $E(\alpha^2)$ (see, formula (\ref{qq}).
406: To investigate whether this is the case for Fe~I we plot the
407: energies of few low odd states of Fe~I with total momentum $J$ = 3, 4 and 5
408: as function of $\alpha^2$ from non-relativistic limit $\alpha=0$ to
409: the physical value of $\alpha$. The results are presented on
410: Figs~\ref{f3}, \ref{f4} and \ref{f5}. As can be seen from the pictures,
411: there are multiple level crossing for states with $J=3$ and $J=4$.
412: However, all these crossings take place on safe distance from the
413: physical value of $\alpha$ and are very unlikely to cause
414: the instability of the results. Another interesting thing to note
415: is that the energies are practically linear functions of $\alpha^2$ in
416: all cases.
417:
418: \begin{table}
419: \caption{Energy levels (cm$^{-1}$) and $g$-factors of the lowest
420: even states of Fe}
421: \label{Fe-even}
422: \begin{ruledtabular}
423: \begin{tabular}{l l c r r r r}
424: Conf. & Term & $J$ & \multicolumn{2}{c}{Experiment\footnotemark[1]} &
425: \multicolumn{2}{c}{Calculations} \\
426: & & & \multicolumn{1}{c}{Energy} & \multicolumn{1}{c}{$g$} &
427: \multicolumn{1}{c}{Energy} & \multicolumn{1}{c}{$g$} \\
428: \hline
429: $3d^64s^2$ & $a \ ^5$D & 4 & 0.000 & 1.500 & 0 & 1.4995 \\
430: & & 3 & 415.932 & 1.500 & 464 & 1.4997 \\
431: & & 2 & 704.004 & 1.500 & 790 & 1.4998 \\
432: & & 1 & 888.129 & 1.500 & 1000 & 1.4998 \\
433: & & 0 & 978.072 & & 1103 & 0.0000 \\
434: & & & & & & \\
435: $3d^74s$ & $a \ ^5$F & 5 & 6928.266 & 1.400 & 6862 & 1.3996 \\
436: & & 4 & 7376.760 & 1.350 & 7374 & 1.3496 \\
437: & & 3 & 7728.056 & 1.249 & 7779 & 1.2497 \\
438: & & 2 & 7985.780 & 0.999 & 8078 & 1.0000 \\
439: & & 1 & 8154.710 &-0.014 & 8275 & 0.0010 \\
440: & & & & & & \\
441: $3d^74s$ & $a \ ^3$F & 4 & 11976.234 & 1.254 & 13040 & 1.2496 \\
442: & & 3 & 12560.930 & 1.086 & 13702 & 1.0835 \\
443: & & 2 & 12968.549 & 0.670 & 14171 & 0.6676 \\
444: \end{tabular}
445: \footnotetext[1]{NIST, Ref.~\cite{NIST}}
446: \end{ruledtabular}
447: \end{table}
448:
449: Table~\ref{Fe-even} presents experimental and theoretical energies
450: and $g$-factors of the lowest even states of Fe~I. The $g$-factors are
451: useful for the identification of the states and for control of
452: configuration mixing~\cite{Dzuba02}. As can be seen the
453: experimental data are reproduced in the calculations with very
454: good accuracy for both the $3d^64s^2$ and $3d^74s$ configurations.
455:
456: \begin{table}
457: \caption{Energy levels (cm$^{-1}$), $g$-factors and relativistic
458: energy shifts ($q$-factors, cm$^{-1}$) for the states of
459: of the $3d^64s4p$ configuration of Fe.}
460: \label{Fe-odd1}
461: \begin{ruledtabular}
462: \begin{tabular}{l c r r r r r}
463: Term & $J$ & \multicolumn{2}{c}{Experiment\footnotemark[1]} &
464: \multicolumn{3}{c}{Calculations} \\
465: & & \multicolumn{1}{c}{Energy} & \multicolumn{1}{c}{$g$} &
466: \multicolumn{1}{c}{Energy} & \multicolumn{1}{c}{$g$} &
467: \multicolumn{1}{c}{$q$} \\
468: \hline
469: z $^7$D$^o$ & 5 & 19350.892~ & 1.597 & 19166 & 1.5987 & 490\\
470: & 4 & 19562.440~ & 1.642 & 19390 & 1.6490 & 662\\
471: & 3 & 19757.033~ & 1.746 & 19611 & 1.7485 & 891\\
472: & 2 & 19912.494~ & 2.008 & 19793 & 1.9976 & 1092\\
473: & 1 & 20019.635~ & 2.999 & 19921 & 2.9950 & 1237\\
474: & & & & & & \\
475: z $^7$F$^o$ & 6 & 22650.421~ & 1.498 & 21663 & 1.4997 & 582 \\
476: & 5 & 22845.868~ & 1.498 & 21891 & 1.5005 & 827 \\
477: & 4 & 22996.676~ & 1.493 & 22062 & 1.5026 & 982 \\
478: & 3 & 23110.937~ & 1.513 & 22189 & 1.5029 & 1103 \\
479: & 2 & 23192.497~ & 1.504 & 22282 & 1.5026 & 1184 \\
480: & 1 & 23244.834~ & 1.549 & 22338 & 1.5029 & 1227 \\
481: & 0 & 23270.374~ & & 22366 & 0.0000 & 1246 \\
482: & & & & & & \\
483: z $^7$P$^o$ & 4 & 23711.457~ & 1.747 & 22543 & 1.7470 & 491 \\
484: & 3 & 24180.864~ & 1.908 & 23034 & 1.9136 & 983 \\
485: & 2 & 24506.919~ & 2.333 & 23440 & 2.3309 & 1316 \\
486: & & & & & & \\
487: z $^5$D$^o$ & 4 & 25899.987\footnotemark[2]
488: & 1.502 & 26428 & 1.4979 & 999 \\
489: & 3 & 26140.177~ & 1.500 & 26679 & 1.4984 & 1223 \\
490: & 2 & 26339.691~ & 1.503 & 26924 & 1.4976 & 1450 \\
491: & 1 & 26479.376~ & 1.495 & 27094 & 1.4971 & 1616 \\
492: & 0 & 26550.476~ & & 27174 & 0.0000 & 1705 \\
493: & & & & & & \\
494: z $^5$F$^o$ & 5 & 26874.549\footnotemark[2]
495: & 1.399 & 27432 & 1.3999 & 880 \\
496: & 4 & 27166.819~ & 1.355 & 27702 & 1.3517 & 1180 \\
497: & 3 & 27394.688~ & 1.250 & 27947 & 1.2530 & 1402 \\
498: & 2 & 27559.581~ & 1.004 & 28119 & 1.0041 & 1568 \\
499: & 1 & 27666.346~ &-0.012 & 28213 & 0.0062 & 1680 \\
500: & & & & & & \\
501: z $^5$P$^o$ & 3 & 29056.321\footnotemark[2]
502: & 1.657 & 29340 & 1.6643 & 859 \\
503: & 2 & 29469.020~ & 1.835 & 29795 & 1.8307 & 1310 \\
504: & 1 & 29732.733~ & 2.487 & 30118 & 2.4966 & 1594 \\
505: & & & & & & \\
506: z $^3$F$^o$ & 4 & 31307.243~ & 1.250 & 32356 & 1.2504 & 1267 \\
507: & 3 & 31805.067~ & 1.086 & 32883 & 1.0885 & 1808 \\
508: & 2 & 32133.986~ & 0.682 & 33263 & 0.6767 & 2177 \\
509: & & & & & & \\
510: z $^3$D$^o$ & 3 & 31322.611~ & 1.321 & 32032 & 1.3314 & 1456 \\
511: & 2 & 31686.346~ & 1.168 & 32464 & 1.1662 & 1843 \\
512: & 1 & 31937.316~ & 0.513 & 32750 & 0.5035 & 2119 \\
513: % & & & & & & \\
514: %z $^3$P$^o$ & 2 & 33946.929 & 1.493 & 34803 & 1.4896 & 1906 \\
515: % & 1 & 34362.871 & 1.496 & 35278 & 1.4968 & 1745 \\
516: % & 0 & 34555.60 & & 35493 & 0.0000 & 1945 \\
517: \end{tabular}
518: \footnotetext[1]{NIST, Ref.~\cite{NIST}}
519: \footnotetext[2]{States observed in quasar absorption spectra}
520: \end{ruledtabular}
521: \end{table}
522:
523: Table~\ref{Fe-odd1} presents experimental and theoretical energies
524: and $g$-factors of the lowest odd states of Fe~I in which the
525: $3d^64s4p$ configuration dominates. Theoretical relativistic
526: frequency shifts ($q$-coefficients) are also presented. The
527: $q$-coefficients were obtained by numerical differentiation
528: using formula~(\ref{deriv05}). Note that only states with
529: $J$=3,4 and 5, for which electric dipole transition to the ground
530: state is possible are needed for the analysis. However, we
531: present $q$-coefficients for all states for better illustration of
532: the accuracy of the calculations. In the linear in $\alpha^2$
533: approximation the difference in $q$-coefficients for states of the
534: same fine-structure multiplet is equal to the fine structure
535: interval between this states. As can be seen from Figs.~\ref{f3},\ref{f4}
536: and \ref{f5} the dependence of the energies on $\alpha^2$ is very
537: close to linear indeed. Therefore, comparing the data for the fine
538: structure and $q$ is another test of the calculations.
539:
540: \begin{table}
541: \caption{Energy levels (cm$^{-1}$), $g$-factors and relativistic
542: energy shifts ($q$-factors, cm$^{-1}$) for the states of the
543: $3d^74p$ and $3d^54s^24p$ configurations of Fe}
544: \label{Fe-odd2}
545: \begin{ruledtabular}
546: \begin{tabular}{l l c r r r r r}
547: Conf. & Term & $J$ & \multicolumn{2}{c}{Experiment\footnotemark[1]} &
548: \multicolumn{3}{c}{Calculations} \\
549: & & & \multicolumn{1}{c}{Energy} & \multicolumn{1}{c}{$g$} &
550: \multicolumn{1}{c}{Energy} & \multicolumn{1}{c}{$g$} &
551: \multicolumn{1}{c}{$q$} \\
552: \hline
553: $3d^74p$ & y $^5$D$^o$ & 4 & 33095.937\footnotemark[2]
554: & 1.496 & 32680 & 1.4511 & 2494 \\
555: & & 3 & 33507.120\footnotemark[2]
556: & 1.492 & 33134 & 1.3492 & 3019 \\
557: & & 2 & 33801.567~ & 1.495 & 33466 & 1.1053 & 3423 \\
558: & & 1 & 34017.098~ & 1.492 & 33705 & 0.1658 & 3754 \\
559: & & 0 & 34121.58~~ & & 34007 & 0.0000 & 3723 \\
560: & & & & & & & \\
561: $3d^74p$ & y $^5$F$^o$ & 5 & 33695.394\footnotemark[2]
562: & 1.417 & 32522 & 1.3964 & 2672 \\
563: & & 4 & 34039.513~ & 1.344 & 33029 & 1.3913 & 3021 \\
564: & & 3 & 34328.749~ & 1.244 & 33404 & 1.3881 & 3317 \\
565: & & 2 & 34547.206~ & 0.998 & 33705 & 1.3777 & 3536 \\
566: & & 1 & 34692.144~ &-0.016 & 33909 & 1.3347 & 3678 \\
567: & & & & & & & \\
568: $3d^74p$ & z $^5$G$^o$ & 5 & 34782.416~ & 1.218 & 33978 & 1.2487 & 3024 \\
569: & & 6 & 34843.94~~ & 1.332 & 33687 & 1.3330 & \\
570: & & 4 & 35257.319~ & 1.103 & 34363 & 1.1406 & 3520 \\
571: & & 3 & 35611.619~ & 0.887 & 34661 & 0.9153 & 3864 \\
572: & & 2 & 35856.400~ & 0.335 & 34883 & 0.3505 & 3464 \\
573: & & & & & & & \\
574: $3d^74p$ & z $^3$G$^o$ & 5 & 35379.206~ & 1.248 & 34506 & 1.2209 & 3340 \\
575: & & 4 & 35767.561~ & 1.100 & 35042 & 1.0731 & 3697 \\
576: & & 3 & 36079.366~ & 0.791 & 35474 & 0.7671 & 4096 \\
577: & & & & & & & \\
578: $3d^74p$ & y $^3$F$^o$ & 4 & 36686.164~ & 1.246 & 35697 & 1.2425 & 3085 \\
579: & & 3 & 37162.740~ & 1.086 & 36227 & 1.0863 & 3487 \\
580: & & 2 & 37521.157~ & 0.688 & & & \\
581: $3d^54s^24p$ & y $^7$P$^o$ & 2 & 40052.030~ & 2.340 & 40529 & 2.3278 & \\
582: & & 3 & 40207.086~ & 1.908 & 40677 & 1.8883 & -2472 \\
583: & & 4 & 40421.85~~ & 1.75~ & 40926 & 1.7491 & -2287 \\
584: \end{tabular}
585: \footnotetext[1]{NIST, Ref.~\cite{NIST}}
586: \footnotetext[2]{States observed in quasar absorption spectra}
587: \end{ruledtabular}
588: \end{table}
589:
590: Table~\ref{Fe-odd2} presents the data similar to those of Table~\ref{Fe-odd1}
591: but for the states where the $3d^74p$ and $3d^54s^24p$ configurations
592: dominate. The values of the $q$-coefficients for the states of the
593: $3d^74p$ configuration are larger than those of the $3d^64s4p$ configuration.
594: This is due to additional contribution from the $4s - 3d$ singe-electron
595: transition as it was explained above.
596:
597: It is interesting that similar to the case of the ion Fe~II~\cite{Dzuba02}
598: neutral iron also has some negative shifters($q<0$). Corresponding states
599: belong to the $3d^54s^24p$ configuration. Negative sign of $q$ is due to
600: the dominant contribution from the $4p -3d$ single-electron transition.
601: The data are presented in Table~\ref{Fe-odd2}. Note however that the spin
602: of these states is different from the spin in the ground state. This means
603: that the electric dipole transition is suppressed by conservation of spin
604: and goes only due to relativistic effects. This in turn probably means
605: that the transitions may be too weak to be observed.
606:
607: We estimate the accuracy of present calculations of the $q$-coefficients
608: to be on the level of 20 to 30\%. The results were obtained with a very
609: simple method which uses small number of basis functions and some
610: semi-empirical fitting. The main challenges for more accurate
611: calculations are strong configuration mixing and large number of
612: valence electrons.
613: Further development of the methods or the use of supercomputers
614: might be needed for better accuracy of the calculations.
615:
616: \section{Conclusion}
617:
618: We have calculated relativistic frequency shifts for a number of the
619: lower odd states of Fe~I. Some of these states were observed in the
620: quasar absorption spectra. Calculations show that due to strong
621: configuration mixing the values of the shifts are large and vary
622: significantly between the states. This makes Fe~I to be a good
623: candidate for the search of variation of the fine structure
624: constant in quasar absorption spectra.
625:
626: \section*{Acknowledgments}
627:
628: We are grateful to Prof. P. Molaro for brining to our attention
629: lines of Fe observed in quasar absorption spectra.
630: The work was funded in part by the Australian Research Council.
631:
632:
633: \begin{thebibliography}{999}
634:
635: \frenchspacing
636:
637: \bibitem{Uzan}
638: J-P. Uzan, Rev.\ Mod.\ Phys. {\bf 75}, 403 (2003).
639:
640: \bibitem{Flambaum07a} V. V. Flambaum,
641: Int. J. Mod. Phys. A {\bf 22}, 4937 (2007).
642:
643: \bibitem{Webb99} J. K. Webb, V. V. Flambaum, C. W. Churchill,
644: M. J. Drinkwater, and J. D. Barrow, Phys.\ Rev.\ Lett. {\bf 82}, 884 (1999).
645:
646: \bibitem{Webb01} J. K. Webb, M. T. Murphy, V. V. Flambaum,
647: V. A. Dzuba, J. D. Barrow, C. W. Churchill, J. X. Prochaska, and
648: A. M. Wolfe, Phys.\ Rev.\ Lett. {\bf 87}, 091301 (2001).
649:
650: \bibitem{Murphy01a} M. T. Murphy, J. K. Webb, V. V. Flambaum,
651: V. A. Dzuba, C. W. Churchill, J. X. Prochaska, J. D. Barrow, and
652: A. M. Wolfe, Not.\ R. Astron.\ Soc. {\bf 327}, 1208 (2001).
653:
654: \bibitem{Murphy01b} M. T. Murphy, J. K. Webb, V. V. Flambaum,
655: C. W. Churchill, and J. X. Prochaska,
656: Not.\ R. Astron.\ Soc. {\bf 327}, 1223 (2001).
657:
658: \bibitem{Murphy01c} M. T. Murphy, J. K. Webb, V. V. Flambaum,
659: C. W. Churchill, and J. X. Prochaska,
660: Not.\ R. Astron.\ Soc. {\bf 327}, 1237 (2001).
661:
662: \bibitem{Murphy01d} M. T. Murphy, J. K. Webb, V. V. Flambaum,
663: M. J. Drinkwater, F. Combes, and T. Wiklind,
664: Not.\ R. Astron.\ Soc. {\bf 327}, 1244 (2001).
665:
666: \bibitem{vlt1} R. Quast, D. Reimers, and S. A. Levshakov,
667: Astron.\ Astrophys. {\bf 417}, L7 (2004).
668:
669: \bibitem{vlt2} R. Srianand, H. Chand, P. Petitjean, and
670: B. Aracil, Astron.\ Astrophys. {\bf 417}, 853 (2004);
671: Phys.\ Rev.\ Lett. {\bf 92}, 121302 (2004).
672:
673: \bibitem{Flambaum07} M. T. Murphy, J. K. Webb, and V. V. Flambaum,
674: arXiv:astro-ph/0612407 (2006); arXiv:astro-ph/0611080 (2006);
675: arXiv:0708.3677 (2007).
676:
677: \bibitem{Srianand07} R. Srianand, H. Chand, P. Petitjean, and B. Aracil ,
678: arXiv:0711.1742 (2007).
679:
680: \bibitem{Dzuba99} V. A. Dzuba, V. V. Flambaum, and J.K. Webb,
681: % Space-Time Variation of Physical Constants and Relativistic
682: % Corrections in Atoms,
683: Phys.\ Rev.\ Lett., {\bf 82}, 888 (1999).
684:
685:
686: \bibitem{Dzuba99a} V. A. Dzuba, V. V. Flambaum, J.K. Webb,
687: % Calculations of the Relativistic Effects in Many-Electron Atoms and
688: % Space-Time Variation of Fundamental Constants,
689: Phys.\ Rev.\ A{\bf 59}, 230 (1999).
690:
691: \bibitem{Dzuba01} V. A. Dzuba, V. V. Flambaum, M. T. Murphy and J. K. Webb,
692: % Relativistic effects in Ni~II and
693: % search for variation of the fine-structure constant,
694: Phys.\ Rev.\ A{\bf 63}, 042509 (2001).
695:
696: \bibitem{Dzuba02} V. A. Dzuba, V. V. Flambaum, M. G. Kozlov, and M. Marchenko,
697: % Alpha dependence of transition frequencies for ions SiII, CrII, FeII,
698: % NiII, and ZnII,
699: Phys.\ Rev.\ A{\bf 66}, 022501 (2002).
700:
701: \bibitem{Berengut04} J. C. Berengut, V. A. Dzuba, V. V. Flambaum,
702: and M. V. Marchenko,
703: % Alpha dependence of transition frequencies for some ions of Ti, Mn, Na,
704: % C, and O and the search for variation of the fine-structure constant,
705: % Na I and C IV
706: Phys. Rev. A{\bf 70}, 064101 (2004).
707:
708: \bibitem{Berengut05} J. C. Berengut, V. V. Flambaum, and M. G. Kozlov ,
709: % Improved calculation of relativistic shift and isotope shift in Mg I
710: Phys.\ Rev.\ A{\bf 72} 044501, (2005).
711:
712: \bibitem{Berengut06} J. C. Berengut, V. V. Flambaum, and M. G. Kozlov,
713: % Calculation of isotope shifts and relativistic shifts in CI, CII, CIII and CIV
714: Phys.\ Rev.\ A{\bf 73} 012504, (2006).
715:
716: \bibitem{archDzuba} J. C. Berengut, V. A. Dzuba, V. V. Flambaum, M. V.
717: Marchenko, J.K. Webb, M. G. Kozlov, and M. T. Murphy,
718: arXiv:physics/0408017 (2006).
719:
720: \bibitem{Porsev} S.G. Porsev, K.V. Koshelev, I.I. Tupitsyn, M.G. Kozlov,
721: D. Reimers, and S.A. Levshakov,
722: % Transition frequency shifts with fine-structure constant variation for Fe II:
723: % Breit and core-valence correlation correction
724: preprint: arXiv:0708.1662 (2007).
725:
726: \bibitem{Johnson} V. A. Dzuba, and W. R. Johnson, arXiv:physics/0710.3417 (2007).
727:
728: \bibitem{Savukov} I. Savukov, and V.A. Dzuba, arXiv:physics/0710.4878 (2007).
729:
730: \bibitem{Molaro} P. Molaro, private communication, (2007).
731:
732: \bibitem{Grant} I. P. Grant, Comput. Phys. Commun. {\bf 84}, 59 (1994).
733:
734: \bibitem{JETP} V. A. Dzuba, V. V. Flambaum, and M.G. Kozlov,
735: JETP Lett. {\bf 63} 882, (1996).
736:
737: \bibitem{Kozlov96} V. A. Dzuba, V. V. Flambaum, and M.G. Kozlov,
738: Phys.\ Rev.\ A{\bf 54} 3948, (1996).
739:
740: \bibitem{Johnson98} V.A. Dzuba, and W.R. Johnson,
741: Phys.\ Rev.\ A{\bf 57} 2459, (1998).
742:
743: \bibitem{VN} V. A. Dzuba,
744: % $V^{N-M}$ approximation for atomic calculations,
745: Phys. Rev. A, {\bf 71}, 032512 (2005).
746:
747: \bibitem{VN1} V. A. Dzuba and V. V. Flambaum,
748: % Core-valence correlations for atoms with open shells,
749: Phys. Rev. A. {\bf 75}, 052504 (2007).
750:
751:
752: \bibitem{NIST} Yu.\ Ralchenko, F.-C. Jou, D.E. Kelleher, A. E. Kramida,
753: A. Musgrove,
754: J. Reader, W.L. Wiese, and K. Olsen (2007).
755: NIST Atomic Spectra Database (version 3.1.3),
756: [Online]. Available: http://physics.nist.gov/asd3
757: [2007, September 18].
758: National Institute of Standards and Technology, Gaithersburg, MD.
759:
760: \end{thebibliography}
761:
762: \end{document}
763: