Brought to you by:

s-PROCESSING IN THE GALACTIC DISK. I. SUPER-SOLAR ABUNDANCES OF Y, Zr, La, AND Ce IN YOUNG OPEN CLUSTERS

, , , , and

Published 2011 July 15 © 2011. The American Astronomical Society. All rights reserved.
, , Citation E. Maiorca et al 2011 ApJ 736 120 DOI 10.1088/0004-637X/736/2/120

0004-637X/736/2/120

ABSTRACT

In a recent study based on homogeneous barium abundance measurements in open clusters (OCs), a trend of increasing [Ba/Fe] ratios for decreasing cluster age was reported. We present here further abundance determinations, relative to four other elements having important s-process contributions, with the aim of investigating whether or not the growth found for [Ba/Fe] is indicative of a general property, shared also by the other heavy elements formed by slow neutron captures. In particular, we derived abundances for yttrium, zirconium, lanthanum, and cerium, using equivalent width measurements and the MOOG code. Our sample includes 19 OCs of different ages, for which the spectra were obtained by the ESO Very Large Telescope using the UVES spectrometer. The growth previously suggested for Ba is confirmed for all the elements analyzed in our study. This fact implies significant changes in our views of the Galactic chemical evolution for elements beyond iron. Our results necessarily require that very low mass asymptotic giant branch stars (M⪅1.5 M) produce larger amounts of s-process elements (and hence activate the 13C-neutron source more effectively) than previously expected. Their role in producing neutron-rich elements in the Galactic disk has been so far underestimated, and their evolution and neutron-capture nucleosynthesis should now be reconsidered.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

This paper presents photospheric abundance determinations of four elements synthesized primarily by slow neutron captures in a large sample of Galactic open clusters (OCs). Specifically, we measured the abundance of yttrium and zirconium, which belong to the first peak of the heavy-element abundance distribution beyond iron, near the neutron-magic number N = 50 (often called light s-nuclei, or ls) along with the abundance of lanthanum and cerium, which are close to the second peak, at N = 82 (s-process isotopes in this region are often called heavy s-nuclei, or hs). The elements we consider here (not only certain isotopes in them) are usually called s-elements. This is certainly conditioned by what we know of the solar composition and does not necessarily apply to the Galaxy as a whole. As an example, it is known that the Ba abundance observed in the Galactic halo is primarily synthesized by the r-process. However, as we deal with the Galactic disk, and as our results suggest that in OCs the s-process component is even more important than in the Sun, we shall consider for our purposes the usual definition of s-elements as appropriate to Y, Zr, La, and Ce.

The clusters in our sample cover a wide range of ages and Galactocentric distances, which allows us to reconstruct the evolution of s-element abundances from the formation of the Galactic disk up to rather recent epochs. This should be compared with the predictions of Galactic chemical evolution models, which in turn require yields from nucleosynthesis computations in stellar models. This topic is therefore closely linked to the modeling of low-mass stars (LMSs), as they are the main contributors to the production of s-elements above A ≃ 85 (the so-called main component of the s-process) during their thermally pulsing asymptotic giant branch (AGB) phases (see, e.g., Busso et al. 1999).

Recently, D'Orazi et al. (2009) measured barium abundance in a large sample of OCs, finding for the first time the surprising result that barium increases with respect to iron as cluster age decreases. This is at odds with expectations of previous Galactic chemical evolution models based on the s-process scenario consolidated in the 1990s (see, e.g., Raiteri et al. 1999; Travaglio et al. 1999). The results by D'Orazi et al. (2009) were based on relatively unevolved stars, which have inherited Ba from the previous evolution of the Galaxy rather than producing it themselves. They showed that chemical evolution models could reproduce the observed trend only by introducing barium yields from very low mass stars enhanced with respect to estimates present in the literature (see Busso et al. 2001). These old estimates are adequate to reproduce the observed abundance trend only up to the solar formation (Travaglio et al. 1999). In fact, D'Orazi et al. (2009) ascertained that they needed a larger production of Ba only from stars with masses between 1 and 1.5 M, which did not contribute significantly in determining the solar composition. Similar results were subsequently obtained by Jacobson et al. (2011a).

We refer to Busso et al. (1999) for a detailed description of the s-element nucleosynthesis in stars. Recent updates are in Straniero et al. (2009) and Cristallo et al. (2009). We briefly recall here that LMSs, during AGB phases, produce s-elements mainly through the activation of the 13C(α, n)16O reaction (the so-called 13C neutron source). Producing 13C in the He intershell during the interpulse periods requires the injection of protons from the envelope into layers below those fully mixed by the third dredge-up, through some form of extra-mixing. Afterward, the previously formed 12C nuclei can undergo proton captures creating 13C seeds. Heavy nuclei near the valley of β-stability present in the He intershell, starting from the same Fe isotopes, can thus capture neutrons, activating s-processing. Gallino et al. (1998) showed that the efficiency of the s-process depends on metallicity, while in those years no relationship with the stellar mass (in the range 1.5–3 M) could be clearly ascertained. The results by D'Orazi et al. (2009) for Ba now seem to integrate this general view, suggesting that the efficiency of the s-process is larger in smaller masses (M ⩽ 1.5 M).

If this anticorrelation between the stellar mass and the effectiveness of neutron-capture nucleosynthesis exists, then one would expect the abundances of other s-elements to be affected in a way similar to Ba. In order to verify the above suggestions and to put solid constraints on s-process nucleosynthesis in the recent Galaxy, we derived the abundances of the four neutron-capture elements mentioned above (Y, Zr, La, and Ce) in a sample of stars from OCs, similar to the one analyzed by D'Orazi et al. (2009).

Our paper is structured as follows: In Section 2 we describe the OC sample along with the methods of analysis used to derive our abundances. In Section 3 we present our main results. A preliminary discussion of the expected consequences is presented in Section 4, and some conclusions are drawn in Section 5.

2. SAMPLE AND ANALYSIS

2.1. Open Cluster Sample

The sample includes 19 OCs with ages in the range of ∼0.1–8.4 Gyr and Galactocentric distances (RGC) between 7.05 and 22 kpc. This sample overlaps considerably with the one from D'Orazi et al. (2009); therefore, we refer to this paper and to the references therein for additional information on the clusters in common and on their spectroscopy. In particular, we adopt here the same cluster ages used by D'Orazi et al. (2009); these are on a homogeneous scale and were taken from Magrini et al. (2009).

A few differences are, however, present in our sample; namely, we do not include the young clusters IC 2602, IC 2391, and NGC 6475, since the available spectral range did not cover useful lines for the abundance analysis. On the other hand, we added three new clusters: IC 4756, NGC 5822 (Pace et al. 2010), and NGC 6192 (Magrini et al. 2010). The spectra of the stars in these clusters were all collected with UVES consistently with the rest of the sample. Details on the observations and instrument configuration can be found in Pace et al. (2010) and Magrini et al. (2010). Also, and most important, the adopted ages for the clusters we study are on the same scale as those of D'Orazi et al. (2009).

The full list of sample clusters is given in Table 1, as are their ages (Column 2), Galactocentric distances (Column 3), metallicities (Column 4), number of stars analyzed, and spectral regions considered. We analyzed spectra of main-sequence (MS) stars for the first eight OCs in the table, and spectra of red giant branch (RGB) or clump stars for the remaining ones, where MS stars are too faint to be observed with UVES.

Table 1. Our OC Sample

Open Cluster Age (Gyr) RGC [Fe/H] References No. of Stars Spectral Region (nm)
NGC 2516 0.11 8.48 +0.01 ± 0.07 1,2 3 480–680
Hyades 0.7 8.50 +0.13 ± 0.05 3,2 4 480–680
Praesepe 0.7 8.62 +0.27 ± 0.04 4 7 330–680
IC 4756 0.79 8.11 +0.01 ± 0.04 5 3 330–450
NGC 5822 1.16 7.87 +0.05 ± 0.03 5 2 330–450
NGC 3680 1.4 8.27 −0.04 ± 0.03 4 2 330–680
IC 4651 1.7 7.70 +0.12 ± 0.05 4 5 330–680
M67 4.3 9.05 +0.03 ± 0.01 6,4 15 330–680
NGC 6192 0.18 7.05 +0.12 ± 0.04 7 4 480–680
NGC 2324 0.7 11.65 −0.17 ± 0.05 8 6 480–680
NGC 3960 0.7 7.96 +0.02 ± 0.04 9 5 480–680
NGC 2660 0.8 9.18 +0.04 ± 0.04 9 5 480–680
NGC 2477 1.0 8.94 +0.07 ± 0.03 8 6 480–680
Berkeley 29 4.3 22.0 −0.31 ± 0.03 10 4 480–680
Berkeley 20 4.3 16.4 −0.30 ± 0.02 10 2 480–680
NGC 6253 4.5 7.10 +0.39 ± 0.08 11 5 480–680
Melotte 66 5.4 10.21 −0.33 ± 0.03 10 6 480–680
Berkeley 32 6.1 11.35 −0.29 ± 0.04 9 9 480–680
Collinder 261 8.4 7.52 +0.13 ± 0.05 10 7 480–680

Notes. We list: cluster ages (Column 2); the Galactocentric distance (Column 3), which has been computed assuming a solar Galactocentric distance RGC☉  =  8.5 kpc; [Fe/H] values (Column 4); references for the metallicity and stellar parameters (Column 5); the number of stars and wavelength range of the spectra analyzed (Columns 6 and 7). References. (1) Terndrup et al. 2002; (2) Randich et al. 2007; (3) Paulson et al. 2003; (4) Pace et al. 2008; (5) Pace et al. 2010; (6) Randich et al. 2006; (7) Magrini et al. 2010; (8) Bragaglia et al. 2008; (9) Sestito et al. 2006; (10) Sestito et al. 2008; (11) Sestito et al. 2007.

Download table as:  ASCIITypeset image

2.2. Selected Lines

In Table 2 we list the lines of Y, Zr, Ce, and La used in the present analysis, along with the atomic parameters and their references. All the lines were inspected very carefully on the UVES spectrum of the Sun, and, in particular, the presence of possible blends was checked. Note that all lines belong to ionized species. Following previous works in the literature, we did not consider hyperfine structure (HFS) for Ce, Y, and Zr (Lawler et al. 2009; Hannaford et al. 1982; Sneden 1973). On the other hand, La needs HFS calculations, and we adopted those used in Lawler et al. (2001). None of the lines is known to be affected by non-LTE (NLTE) effects.

Table 2. Line List and Solar Abundances Line-by-line

Wavelength EP log  gf EW A(X)
(nm) (eV)   (mÅ)  
Y IIa
412.491 0.41 −1.50 22.8 2.24
439.801 0.13 −1.00 51.5 2.18
485.487 0.99 −0.38 48.1 2.26
488.369 1.08 +0.07 59.1 2.18
490.012 1.03 −0.09 56.6 2.23
508.743 1.08 −0.17 49.4 2.15
520.041 0.99 −0.57 38.2 2.18
Zr IIb
405.032 0.71 −1.00 23.2 2.53
420.898 0.71 −0.46 44.0 2.52
437.974 1.53 −0.25 25.5 2.60
La IIc
398.857 0.40 −0.455 51.3 1.12
399.577 0.17 −0.686 39.0 1.08
408.671 0.00 −0.696 40.5 1.12
433.377 0.17 −0.686 48.1 1.20
Ce IId
399.924 0.29 +0.060 18.8 1.56
404.258 0.49 +0.000 10.6 1.49
407.347 0.48 +0.210 18.8 1.58
408.322 0.70 +0.270 18.0 1.71
413.765 0.52 +0.400 29.8 1.74
414.500 0.70 +0.100 10.0 1.54
434.979 0.70 −0.320 5.4 1.64
518.746 1.21 +0.170 5.3 1.58
527.423 1.04 +0.130 7.3 1.61

Note. References for atomic parameters: (a) Hannaford et al. 1982; (b) Malcheva et al. 2006; (c) Lawler et al. 2001; (d) Lawler et al. 2009.

Download table as:  ASCIITypeset image

As anticipated above, for all the elements used in this analysis, we selected only the spectral features that can be detected in the UVES spectrum of the Sun, in order to be able to derive differential abundances line-by-line with respect to the Sun itself.

Equivalent widths (EWs) of the lines listed in Table 2 were measured using the splot routine within the IRAF package.4 Whenever possible, measurements were performed by means of direct integration of lines instead of fitting with Gauss' or Voigt's profiles.

2.3. Solar Abundances

Also for the Sun, abundances were derived from EW measurements using the MOOG code (Sneden 1973; we adopted the 2009 version) and the grid of one-dimensional local thermal equilibrium model atmospheres by Kurucz (1993). Stark and radiative broadening were treated in the standard way, and for collisional damping we adopted the classical Unsöld approximation. In Columns 4 and 5 of Table 2 we show our EW measurements for the Sun along with abundances in the usual form (A(X) = log[N(X)/N(H)]+12). EWs were measured on the solar spectrum obtained with UVES. As in Randich et al. (2006), we adopted the following Solar parameters: Teff = 5770 K, log g = 4.44, and ξ = 1.1 km s−1.

Initially, we had also derived abundances using lines in the wavelength interval 300–390 nm, but the resulting values are systematically lower (∼0.3 dex) than those from lines with λ  >  390 nm. Since the same behavior is also obtained by performing spectral synthesis (C. Sneden 2010, private communication), we believe that some problems in the opacity calculation arise in this UV region. Therefore, we decided to discard these lines, limiting our analysis to those with wavelengths larger than 390 nm.

The mean solar abundances for the four elements derived from our analysis are shown in Table 3, together with the values derived by Asplund et al. (2009, hereafter AS09) using three-dimensional models. The meteoritic abundances from Lodders et al. (2009) are also indicated.

Table 3. Abundances in the Sun

Element Photosphere AS09 Meteorites
Y 2.20 ± 0.04 2.21 ± 0.04 2.19 ± 0.04
Zr 2.55 ± 0.04 2.58 ± 0.04 2.55 ± 0.04
La 1.13 ± 0.05 1.10 ± 0.04 1.19 ± 0.02
Ce 1.61 ± 0.09 1.58 ± 0.04 1.60 ± 0.02

Download table as:  ASCIITypeset image

Our abundances for Y, Zr, and Ce are in very good agreement with both the meteoritic values and AS09 determinations. On the other hand, the abundance of La derived in this study agrees with the value of AS09, but both are lower than in the meteoritic data. Although the difference is within the uncertainties (see Table 3), it is larger than the differences we found for other elements and indicates that additional studies on the solar lanthanum are required.

2.4. Abundances in OCs

We applied the same method used for the Sun to measure abundances in OC stars. Stellar parameters for the sample stars were taken from the reference papers given in Table 1, while in Table 4 we show their values. We underline that all the stellar parameters of this study were determined spectroscopically. As mentioned, our analysis of OC stars is strictly differential with respect to the Sun; thus, for each line we subtracted from the derived abundance the corresponding solar abundance from the same line, obtaining [X/H]line. [X/H] values for each star were then obtained as the average of [X/H]line values. In this way, the uncertainties of the adopted set of atomic parameters, in particular log gf values, cancel out. The [X/Fe] estimate for each star was finally derived by subtracting the [Fe/H] value of the star itself (rather than the mean metallicity of the cluster) from its [X/H] value.

Table 4. Differential Abundances Star by Star

Cluster Star Teff log g ξ [Fe/H] [Y/H]5087 [Zr/H] [La/H] [Ce/H] [Y/H]all
NGC 2516 CTIO2 5238 4.5 0.89 0.01 +0.14 ± 0.03 (1) ... ... +0.21 ± 0.04 (1) +0.18 ± 0.04 (5)
  CTIO3 5110 4.5 0.85 0.01 +0.16 ± 0.03 (1) ... ... +0.20 ± 0.06 (2) +0.33 ± 0.12 (5)
  DK213 5305 4.5 0.91 0.01 +0.20 ± 0.04 (1) ... ... +0.17 ± 0.03 (1) +0.32 ± 0.10 (5)
HYADES vb42 5308 4.5 0.91 0.10 +0.24 ± 0.04 (1) ... ... +0.17 ± 0.05 (2) +0.30 ± 0.06 (5)
  vb109 5141 4.5 0.86 0.13 +0.27 ± 0.04 (1) ... ... +0.27 ± 0.04 (2) +0.42 ± 0.09 (5)
  vb182 5079 4.5 0.84 0.13 +0.19 ± 0.04 (1) ... ... +0.28 ± 0.07 (2) +0.35 ± 0.09 (5)
  vb187 5339 4.5 0.92 0.11 +0.26 ± 0.04 (1) ... ... +0.22 ± 0.07 (2) +0.38 ± 0.08 (5)
Praesepe KW49 6150 4.5 1.41 0.22 +0.20 ± 0.04 (1) +0.24 ± 0.11 (3) +0.26 ± 0.18 (5) +0.24 ± 0.07 (1) +0.24 ± 0.14 (3)
  KW100 6150 4.3 1.78 0.27 ... ... +0.20 ± 0.12 (2) ... ...
  KW208 6280 4.6 1.52 0.28 +0.23 ± 0.03 (1) +0.28 ± 0.10 (3) +0.32 ± 0.15 (5) ... +0.28 ± 0.05 (4)
  KW326 5800 4.5 1.28 0.29 +0.18 ± 0.03 (1) +0.22 ± 0.06 (3) +0.28 ± 0.10 (5) +0.25 ± 0.13 (6) +0.23 ± 0.03 (5)
  KW368 5720 4.5 1.12 0.26 +0.19 ± 0.05 (1) +0.21 ± 0.06 (3) +0.26 ± 0.06 (5) +0.20 ± 0.10 (8) +0.26 ± 0.05 (5)
  KW392 6250 4.6 1.48 0.35 +0.31 ± 0.03 (1) +0.31 ± 0.09 (3) +0.37 ± 0.08 (5) +0.41 ± 0.10 (6) +0.31 ± 0.06 (5)
  KW418 6150 4.4 1.27 0.24 +0.21 ± 0.04 (1) +0.19 ± 0.08 (3) ... +0.10 ± 0.09 (2) +0.27 ± 0.07 (5)
IC 4756 HER97 6118 4.5 1.29 0.00 ... +0.10 ± 0.03 (3) +0.20 ± 0.05 (5) +0.14 ± 0.05 (3) ...
  HER165 6070 4.5 1.24 0.05 ... +0.11 ± 0.13 (3) +0.23 ± 0.01 (5) +0.23 ± 0.05 (7) +0.15 ± 0.07 (2)
  HER240 6007 4.5 1.21 −0.02 ... +0.08 ± 0.04 (3) +0.16 ± 0.06 (5) +0.14 ± 0.10 (7) +0.10 ± 0.07 (2)
NGC 5822 TATM11003 6160 4.7 1.05 0.02 ... +0.35 ± 0.09 (3) +0.33 ± 0.08 (5) +0.27 ± 0.06 (4) ...
  TATM11014 6273 4.7 1.26 0.07 ... +0.25 ± 0.04 (3) +0.38 ± 0.03 (5) +0.32 ± 0.04 (2) ...
NGC 3680 AHTC1009 6010 4.5 1.16 0.00 +0.09 ± 0.03 (1) +0.03 ± 0.08 (3) +0.17 ± 0.04 (5) +0.20 ± 0.09 (8) +0.07 ± 0.08 (5)
  EGG70 6210 4.5 1.36 −0.07 +0.00 ± 0.03 (1) +0.10 ± 0.08 (3) +0.16 ± 0.07 (5) +0.25 ± 0.11 (7) +0.00 ± 0.04 (5)
IC 4651 AMC1109 6060 4.5 1.20 0.11 +0.16 ± 0.03 (1) +0.19 ± 0.05 (3) +0.22 ± 0.05 (5) +0.13 ± 0.06 (6) +0.16 ± 0.03 (5)
  AMC2207 6050 4.4 1.18 0.13 +0.13 ± 0.03 (1) +0.17 ± 0.05 (2) +0.16 ± 0.11 (5) +0.16 ± 0.05 (4) +0.11 ± 0.02 (5)
  AMC4220 5910 4.6 1.06 0.19 +0.17 ± 0.04 (1) +0.29 ± 0.11 (2) +0.24 ± 0.09 (5) +0.25 ± 0.02 (3) +0.23 ± 0.07 (5)
  AMC4226 5980 4.4 1.19 0.13 +0.13 ± 0.04 (1) +0.23 ± 0.06 (1) +0.20 ± 0.11 (5) +0.09 ± 0.04 (2) +0.10 ± 0.04 (4)
  EGG45 6320 4.4 1.50 0.05 +0.12 ± 0.03 (1) +0.11 ± 0.05 (3) +0.15 ± 0.09 (5) +0.09 ± 0.09 (4) +0.05 ± 0.06 (5)
M67 S969 5800 4.4 1.10 0.01 −0.05 ± 0.04 (1) ... ... +0.03 ± 0.06 (2) +0.00 ± 0.04 (5)
  S988 6153 4.1 1.45 0.03 +0.03 ± 0.04 (1) ... ... +0.08 ± 0.04 (2) +0.04 ± 0.07 (5)
  S994 6151 4.1 1.45 0.00 −0.02 ± 0.05 (1) ... ... −0.05 ± 0.07 (2) +0.03 ± 0.11 (5)
  S995 6210 3.9 1.50 0.05 −0.03 ± 0.04 (1) ... ... +0.04 ± 0.06 (2) +0.04 ± 0.05 (4)
  S998 6223 4.0 1.50 0.07 −0.07 ± 0.06 (1) ... ... +0.07 ± 0.06 (1) +0.00 ± 0.06 (4)
  S1034 6019 4.0 1.50 0.01 +0.06 ± 0.04 (1) ... ... +0.07 ± 0.08 (2) +0.06 ± 0.08 (5)
  S1239 5541 3.8 1.25 0.02 +0.01 ± 0.04 (1) ... ... +0.06 ± 0.04 (2) +0.06 ± 0.04 (4)
  S1252 5938 4.4 1.15 0.05 +0.09 ± 0.05 (1) ... ... +0.11 ± 0.07 (2) +0.07 ± 0.04 (5)
  S1256 5907 4.5 1.10 0.06 +0.03 ± 0.04 (1) ... ... +0.03 ± 0.05 (2) +0.04 ± 0.05 (5)
  S746 5750 4.4 0.84 0.07 ... +0.09 ± 0.02 (3) +0.06 ± 0.08 (5) ... ...
  S1048 5900 4.4 0.94 0.03 ... ... +0.08 ± 0.08 (5) +0.12 ± 0.06 (5) ...
  S1092 6160 4.4 1.42 0.07 ... +0.05 ± 0.03 (3) +0.15 ± 0.04 (5) ... ...
  S1255 5840 4.5 1.05 0.01 ... +0.12 ± 0.04 (3) +0.09 ± 0.04 (5) ... +0.12 ± 0.06 (2)
  S1283 6100 4.4 1.16 0.03 ... +0.06 ± 0.04 (3) +0.08 ± 0.05 (5) ... ...
  S1287 6100 4.4 1.17 −0.04 ... +0.00 ± 0.07 (3) +0.05 ± 0.04 (5) ... ...
NGC 6192 9 5230 2.8 1.95 0.21 +0.36 ± 0.07 (1) ... ... ... +0.52 ± 0.13 (4)
  45 5050 2.6 1.65 0.04 +0.19 ± 0.06 (1) ... ... ... +0.37 ± 0.12 (5)
  96 5050 2.5 2.10 0.08 +0.13 ± 0.06 (1) ... ... ... +0.47 ± 0.25 (4)
  137 4670 1.9 1.80 0.00 −0.01 ± 0.08 (1) ... ... ... +0.37 ± 0.22 (5)
NGC 2324 10488 5100 2.2 1.21 −0.17 +0.06 ± 0.06 (1) ... ... +0.01 ± 0.06 (2) +0.18 ± 0.16 (4)
  10787 5060 2.1 1.23 −0.17 +0.07 ± 0.07 (1) ... ... +0.04 ± 0.08 (2) +0.16 ± 0.11 (4)
  10797 5040 2.2 1.21 −0.17 +0.09 ± 0.07 (1) ... ... +0.06 ± 0.07 (2) +0.23 ± 0.15 (5)
  10828 4750 1.6 1.29 −0.10 +0.08 ± 0.07 (1) ... ... +0.06 ± 0.07 (2) +0.37 ± 0.20 (5)
  30878 5000 2.1 1.22 −0.27 −0.04 ± 0.08 (1) ... ... −0.14 ± 0.08 (2) +0.11 ± 0.13 (4)
  30957 5000 1.8 1.26 −0.17 −0.15 ± 0.09 (1) ... ... −0.18 ± 0.09 (2) +0.16 ± 0.20 (5)
NGC 3960 C4 5050 2.5 1.17 0.07 +0.20 ± 0.08 (1) ... ... +0.14 ± 0.05 (2) +0.33 ± 0.13 (5)
  C5 4870 2.2 1.22 0.00 +0.08 ± 0.07 (1) ... ... +0.07 ± 0.06 (2) +0.25 ± 0.13 (5)
  C6 4950 2.4 1.19 0.02 +0.18 ± 0.07 (1) ... ... +0.13 ± 0.05 (2) +0.29 ± 0.13 (4)
  C8 5040 2.6 1.18 0.00 +0.07 ± 0.07 (1) ... ... +0.11 ± 0.05 (2) +0.21 ± 0.15 (5)
  C9 5000 2.5 1.18 0.02 +0.09 ± 0.08 (1) ... ... +0.16 ± 0.06 (2) +0.23 ± 0.11 (5)
NGC 2660 296 5200 3.0 1.11 0.08 +0.31 ± 0.07 (1) ... ... +0.41 ± 0.07 (2) +0.43 ± 0.10 (4)
  318 5030 2.6 1.16 0.00 +0.01 ± 0.10 (1) ... ... +0.23 ± 0.08 (2) +0.33 ± 0.18 (5)
  542 5060 2.5 1.17 0.03 +0.05 ± 0.10 (1) ... ... +0.14 ± 0.05 (2) +0.31 ± 0.24 (5)
  694 5100 2.8 1.14 0.05 +0.28 ± 0.07 (1) ... ... +0.37 ± 0.07 (2) +0.41 ± 0.15 (4)
  862 5100 2.6 1.16 0.02 +0.30 ± 0.06 (1) ... ... +0.25 ± 0.05 (2) +0.31 ± 0.07 (5)
NGC 2477 36280 4980 2.8 1.14 0.05 +0.32 ± 0.08 (1) ... ... +0.23 ± 0.06 (2) +0.39 ± 0.08 (5)
  36288 4970 2.7 1.15 0.05 +0.31 ± 0.08 (1) ... ... +0.26 ± 0.06 (2) +0.50 ± 0.12 (5)
  36326 5000 2.7 1.15 0.10 +0.37 ± 0.08 (1) ... ... +0.26 ± 0.06 (2) +0.51 ± 0.10 (5)
  36363 4950 2.7 1.16 0.05 +0.26 ± 0.06 (1) ... ... +0.22 ± 0.05 (2) +0.45 ± 0.12 (5)
  36385 5030 2.7 1.15 0.07 +0.10 ± 0.09 (1) ... ... +0.09 ± 0.07 (2) +0.44 ± 0.23 (5)
  36449 4970 2.6 1.16 0.12 +0.33 ± 0.08 (1) ... ... +0.22 ± 0.06 (2) +0.54 ± 0.13 (5)
Berkeley 29 257 4930 2.6 1.18 −0.36 −0.19 ± 0.09 (1) ... ... ... −0.10 ± 0.06 (4)
  398 5020 2.7 1.17 −0.31 +0.25 ± 0.08 (1) ... ... ... +0.26 ± 0.08 (5)
  602 4970 2.4 1.28 −0.34 −0.01 ± 0.09 (1) ... ... −0.18 ± 0.09 (2) +0.09 ± 0.22 (4)
  933 4930 2.3 1.25 −0.30 +0.04 ± 0.09 (1) ... ... −0.09 ± 0.09 (1) +0.21 ± 0.10 (5)
Berkeley 20 1201 4700 2.1 1.35 −0.28 −0.53 ± 0.09 (1) ... ... −0.43 ± 0.08 (2) −0.66 ± 0.12 (5)
  1240 4400 1.7 1.27 −0.31 −0.32 ± 0.09 (1) ... ... −0.19 ± 0.11 (2) −0.06 ± 0.16 (5)
NGC 6253 23498 5630 3.4 1.30 0.32 +0.52 ± 0.09 (1) ... ... ... +0.60 ± 0.15 (5)
  23501 6050 3.8 0.90 0.29 +0.10 ± 0.10 (1) ... ... ... +0.25 ± 0.30 (4)
  24707 4940 3.0 1.16 0.39 +0.11 ± 0.10 (1) ... ... +0.35 ± 0.07 (2) +0.48 ± 0.23 (5)
  69885 6200 3.8 1.27 0.45 +0.51 ± 0.09 (1) ... ... ... +0.40 ± 0.16 (4)
  105495 4450 2.5 1.23 0.49 +0.28 ± 0.07 (1) ... ... +0.38 ± 0.08 (2) +0.67 ± 0.25 (5)
Melotte 66 1346 4750 2.0 1.17 −0.37 −0.30 ± 0.06 (1) ... ... −0.41 ± 0.06 (2) −0.22 ± 0.07 (5)
  1493 4770 2.1 1.20 −0.35 −0.32 ± 0.06 (1) ... ... −0.31 ± 0.05 (2) −0.24 ± 0.08 (5)
  1785 4770 2.0 1.20 −0.30 −0.40 ± 0.08 (1) ... ... −0.40 ± 0.06 (2) −0.25 ± 0.10 (5)
  1865 4717 2.0 1.24 −0.34 ... ... ... −0.23 ± 0.06 (2) +0.20 ± 0.09 (4)
  1884 4750 2.5 1.23 −0.30 −0.20 ± 0.10 (1) ... ... −0.16 ± 0.11 (2) −0.08 ± 0.20 (5)
  2212 4850 2.4 1.25 −0.31 −0.34 ± 0.07 (1) ... ... −0.22 ± 0.07 (2) −0.17 ± 0.12 (5)
Berkeley 32 17 4830 2.2 1.21 −0.31 −0.34 ± 0.07 (1) ... ... −0.36 ± 0.06 (2) −0.22 ± 0.09 (5)
  18 4850 2.3 1.21 −0.27 −0.26 ± 0.07 (1) ... ... −0.21 ± 0.07 (2) −0.15 ± 0.06 (5)
  19 4760 2.3 1.21 −0.35 −0.43 ± 0.08 (1) ... ... −0.33 ± 0.07 (2) −0.27 ± 0.12 (5)
  25 4760 2.4 1.19 −0.20 −0.05 ± 0.09 (1) ... ... −0.21 ± 0.07 (2) −0.01 ± 0.08 (5)
  27 4780 2.4 1.19 −0.24 −0.40 ± 0.09 (1) ... ... −0.23 ± 0.09 (2) −0.19 ± 0.19 (5)
  45 4920 3.0 1.11 −0.35 −0.56 ± 0.10 (1) ... ... ... −0.30 ± 0.15 (5)
  938 4870 2.3 1.19 −0.30 −0.46 ± 0.09 (1) ... ... −0.35 ± 0.06 (2) −0.28 ± 0.12 (5)
  940 4800 2.1 1.23 −0.33 −0.40 ± 0.09 (1) ... ... −0.37 ± 0.10 (2) −0.18 ± 0.13 (5)
  941 4760 2.4 1.19 −0.29 −0.10 ± 0.07 (1) ... ... −0.07 ± 0.06 (2) +0.16 ± 0.17 (5)
Collinder 261 2 4350 1.7 1.25 0.12 −0.11 ± 0.07 (1) ... ... ... +0.53 ± 0.39 (5)
  5 4600 2.0 1.24 0.14 −0.18 ± 0.09 (1) ... ... −0.08 ± 0.06 (2) +0.14 ± 0.19 (5)
  6 4500 2.3 1.18 0.16 +0.01 ± 0.07 (1) ... ... ... +0.43 ± 0.26 (5)
  7 4546 2.1 1.20 0.18 −0.03 ± 0.09 (1) ... ... ... +0.43 ± 0.32 (5)
  9 4720 2.0 1.27 0.04 −0.09 ± 0.07 (1) ... ... −0.09 ± 0.07 (2) +0.05 ± 0.14 (5)
  10 4700 2.4 1.20 0.20 +0.06 ± 0.09 (1) ... ... ... +0.32 ± 0.23 (5)
  11 4670 2.1 1.13 0.09 −0.17 ± 0.09 (1) ... ... −0.05 ± 0.08 (2) +0.15 ± 0.19 (5)

Notes. Values in brackets refer to the number of lines used to get the corresponding abundance. Column 7 shows yttrium abundances from the 508.742 nm line only. Column 11 shows yttrium abundances derived through all available lines.

Download table as:  ASCIITypeset images: 1 2

2.5. Errors

Random errors include the contributions from errors of EW measurements and from uncertainties of stellar parameters.

Errors due to EW measurements in general correspond to the line-to-line abundance scatter for a given element, since the possible scatter due to uncertainties in log gf values is minimized by the use of differential analysis with respect to the Sun. However, when fewer than three lines were available, we obtained five EW measurements and computed the corresponding average [El/H] values; the standard deviation of the measurements was then adopted as the error in [El/H].

As is usually done, errors due to uncertainties in stellar parameters were estimated by varying each parameter separately while leaving the others unchanged. The results are shown in Table 5; we performed this analysis for four stars that fairly well cover the parameter space of the whole sample. Uncertainties in effective temperature, surface gravity, and microturbulence were estimated following the discussions in Sestito et al. (2006) and Randich et al. (2006). We found that variations of 100 K in Teff neither produce significant changes in the abundances, nor in red giants, nor in MS stars. On the other hand, and as is expected from an analysis based on lines from ionized elements, a 0.25 dex difference in log  g leads to a variation of ±0.1 dex in the [X/H] value. Finally, changes of ±0.15 km s−1 in the microturbulence parameter ξ of red giants have a rather large effect (about 0.15 dex) for Y, while they are less significant (∼0.05 dex) for Ce (note that we do not have measurements for Zr and La in red giants). The effect of a different ξ is smaller for dwarf stars.

Table 5. Error Estimates Due to Uncertainties in Stellar Parameters

Δ[X/H] ΔTeff = ±100 K Δlog g = ±0.25 dex Δξ = ±0.15 km s−1 Reference
        Star
Δ[Y/H] 0.00/−0.01 0.08/−0.11 −0.15/0.15 (a)
Δ[Ce/H] 0.01/0.00 0.11/−0.11 −0.03/0.06 (a)
Δ[Y/H] 0.00/−0.02 0.08/−0.11 −0.16/0.15 (b)
Δ[Ce/H] 0.01/−0.02 0.10/−0.11 −0.05/0.06 (b)
Δ[Y/H] 0.02/−0.02 0.10/−0.10 −0.06/0.07 (c)
Δ[Ce/H] 0.02/−0.02 0.10/−0.10 0.00/0.01 (c)
Δ[La/H] 0.03/−0.02 0.11/−0.09 −0.01/0.02 (d)
Δ[Zr/H] 0.02/−0.02 0.10/−0.10 −0.02/0.03 (d)

Note. (a) Star 18 of Be32: Teff = 4850 K, log  g = 2.3, and ξ = 1.21 km s−1; (b) Star 10488 of NGC 2324: Teff = 5100 K, log  g = 2.2, and ξ = 1.21 km s−1; (c) Star S994 of M67: Teff = 6151 K, log  g = 4.1, and ξ = 1.45 km s−1; (d) Star sand1092 of M67: Teff = 6160 K, log  g = 4.41, and ξ = 1.42 km s−1.

Download table as:  ASCIITypeset image

Systematic errors might be due to different effects such as the spectrocode, model atmospheres, or estimates of the continuum in the spectra, and of log gf values. While, as mentioned, the effect of the latter should be minimal due to the use of the differential analysis, the other three effects are more difficult to estimate. For three of the sample clusters, measurements of one or more of the elements analyzed in this paper are available in the literature, namely M67 (Tautvaisiene et al. 2000; Yong et al. 2005; Pancino et al. 2010; Önehag et al. 2011), Be 29 (Yong et al. 2005), the Hyades and Collinder 261 (De Silva et al. 2007), and IC 4651 (Mikolaitis et al. 2011). For one cluster only, M67, more than one s-process determination is available in the literature. In Figure 1 we compare our abundances for this cluster with three literature values. The figure indicates that our estimates generally agree within the errors and within ±0.05 dex with previous determinations; however, the two Zr measurements from the literature are considerably lower than our own value. We believe that the latter is likely more correct, since the significantly undersolar [Zr/Fe] derived by Tautvaisiene et al. (2000) and Yong et al. (2005) would not be consistent with the close-to-solar abundance ratios for Y, La, and Ce found by us and by other authors, and, in the more general case, with the solar composition of all the other elements. In any case, possible systematic errors would not affect our results, which are based on the comparison of clusters analyzed with the same method, and thus on the same abundance scale.

Figure 1.

Figure 1. Comparison of our [X/Fe] determinations for M67 with data available in the literature. Different symbols denote different elements, while different colors indicate different studies. Namely: filled circles, open circles, open triangles, and open squares represent Y, Zr, La, and Ce, respectively; black, red, and blue denote Pancino et al. (2010), Yong et al. (2005), and Tautvaisiene et al. (2000), respectively. The dotted lines represent the [Fe/H]our = [Fe/H]lit. ± 0.05 relation.

Standard image High-resolution image

2.5.1. Gravity Uncertainty

In this study we analyzed singly ionized lines for all the elements. Therefore gravity is important in our determinations, as can be seen from Table 5. In Table 4 one can notice a peculiarity. For some MS stars the log g values (which were derived from spectroscopic measurements available in the current literatures) are larger than expected on the basis of their temperature and evolutionary status. As typical examples, we quote KW 208 and KW 392 in the Praesepe cluster, with estimates for log g of 4.6 dex, and the two stars belonging to NGC 5822 with log g  =  4.7 dex. For all of them we would rather select log g values below 4.5, hence we decided to rederive their parameters spectroscopically ourselves. We used the iron line list already adopted in Randich et al. (2006). EWs were measured with daospec, while abundances were obtained with MOOG (2009 version). Our results are summarized in Table 6, where we also list the number of Fe i and Fe ii lines employed. These produced log g values smaller by 0.2–0.3 dex than those of Pace et al. (2008) for KW 208 and KW 392. Using our estimates, the [El/H] abundances for these stars are reduced by less than ∼0.1 dex, according to what was already shown in Table 5. However, the average abundances for Praesepe change only by ∼0.05 dex with respect to those derived with the parameters measured by Pace et al. (2008). Concerning the two stars belonging to NGC 5822, we found a log g reduction of ∼0.1 dex for TATM 11003 and of ∼0.3 dex for TATM 11014. The subsequent cluster mean abundance is less than ∼0.1 dex smaller than the value derived using stellar parameters from the literature. Therefore, these few peculiar cases have a vanishingly small effect on our general trends, so that we can conclude that they do not affect in any significant way our con- clusions.

Table 6. Comparison of Stellar Parameters

Star Name Stellar Parameter Photometry–Literature Spectroscopy–Literature Spectroscopy–This Study
KW208 Teff 5993 6208 6150
  log g 4.45 4.58 4.44
  ξ ... 1.52 1.43
  ♯ Fe i ... ... 31
  ♯ Fe ii ... ... 6
  [Y/H] ... 0.28 ± 0.05 0.20 ± 0.05
KW392 Teff 5902 6250 6070
  log g 4.46 4.56 4.28
  ξ   1.48 1.38
  ♯ Fe i ... ... 36
  ♯ Fe ii ... ... 8
  [Y/H] ... 0.31 ± 0.06 0.21 ± 0.05
TATM11003 Teff 5625 6160 6210
  log g   4.74 4.59
  ξ   1.05 1.08
  ♯ Fe i ... ... 30
  ♯ Fe ii ... ... 7
  [Ce/H] ... 0.27 ± 0.06 0.24 ± 0.05
TATM11014 Teff 5764 6273 6120
  log g   4.74 4.39
  ξ   1.26 1.34
  ♯ Fe i ... ... 28
  ♯ Fe ii ... ... 4
  [Ce/H] ... 0.32 ± 0.04 0.22 ± 0.04

Note. Literature data are from Pace et al. (2008) for KW 208 and KW 392 and from Pace et al. (2010) and references therein for TATM 11003 and TATM 11014.

Download table as:  ASCIITypeset image

3. RESULTS

Our results are summarized in Tables 4 and 7, where we list the average [X/H] and [X/Fe] values for each of the target clusters and the [X/H] values for all the stars analyzed, respectively. As discussed in Section 2.5, the errors listed in Table 4 refer to the line-to-line scatter; on the other hand, errors in Table 7 provide the standard deviation with respect to the cluster mean. We mention in passing that the mean values for Be 32 were computed excluding the star 941, whose abundances deviate by more than 2σ from the average. This star is discussed below in Section 4.2. Tables 4 and 7 put in evidence four major points that will be discussed in detail in the following sections. First, all the sample clusters show a rather homogeneous composition in the four s-process elements, although one star is present in Be 32 with abundances enriched with respect to the mean of the cluster to which it belongs. Second, clusters that appear enriched in first-peak elements (Y and Zr) are also enriched in second-peak elements (La and Ce). Third, clusters where dwarf members were analyzed show a similar enrichment in Y and Ce, while those where the abundances were based on the spectra of giants, together with Hyades stars (dwarfs stars), show higher enrichments in Y. Fourth, and most important, young clusters are generally characterized by higher s-element abundances than are old clusters.

Table 7. Mean Abundances in OCs

Cluster [Y/H] [Y/Fe] [Zr/H] [Zr/Fe] [La/H] [La/Fe] [Ce/H] [Ce/Fe]
NGC 2516 0.17 ± 0.03 0.16 ± 0.03 ... ... ... ... 0.19 ± 0.02 0.18 ± 0.02
Hyades 0.24 ± 0.04 0.12 ± 0.04 ... ... ... ... 0.24 ± 0.05 0.12 ± 0.04
Praesepe 0.27 ± 0.03 −0.01 ± 0.03 0.24 ± 0.05 −0.03 ± 0.03 0.28 ± 0.06 0.00 ± 0.04 0.24 ± 0.11 −0.03 ± 0.08
IC 4756 0.13 ± 0.04 0.11 ± 0.01 0.10 ± 0.02 0.09 ± 0.02 0.20 ± 0.04 0.19 ± 0.01 0.17 ± 0.05 0.16 ± 0.02
NGC 5822 ... ... 0.30 ± 0.07 0.26 ± 0.11 0.36 ± 0.04 0.31 ± 0.01 0.30 ± 0.04 0.25 ± 0.01
NGC 3680 0.03 ± 0.06 0.07 ± 0.01 0.07 ± 0.05 0.10 ± 0.10 0.17 ± 0.01 0.20 ± 0.04 0.23 ± 0.04 0.26 ± 0.08
IC 4651 0.13 ± 0.07 0.01 ± 0.04 0.20 ± 0.07 0.08 ± 0.03 0.19 ± 0.04 0.07 ± 0.03 0.14 ± 0.07 0.02 ± 0.04
M67 0.05 ± 0.04 0.01 ± 0.05 0.06 ± 0.05 0.04 ± 0.05 0.08 ± 0.04 0.06 ± 0.04 0.06 ± 0.05 0.02 ± 0.04
NGC 6192 0.17 ± 0.15 0.08 ± 0.08 ... ... ... ... ... ...
NGC 2324 0.02 ± 0.09 0.19 ± 0.09 ... ... ... ... −0.03 ± 0.11 0.15 ± 0.09
NGC 3960 0.12 ± 0.06 0.10 ± 0.04 ... ... ... ... 0.12 ± 0.03 0.10 ± 0.03
NGC 2660 0.19 ± 0.15 0.15 ± 0.13 ... ... ... ... 0.28 ± 0.11 0.24 ± 0.09
NGC 2477 0.28 ± 0.10 0.21 ± 0.09 ... ... ... ... 0.21 ± 0.06 0.14 ± 0.07
Berkeley 29 0.02 ± 0.18 0.35 ± 0.16 ... ... ... ... −0.14 ± 0.06 0.19 ± 0.04
Berkeley 20 −0.43 ± 0.15 −0.13 ± 0.17 ... ... ... ... −0.31 ± 0.17 −0.01 ± 0.19
NGC 6253 0.30 ± 0.21 −0.08 ± 0.20 ... ... ... ... 0.37 ± 0.02 −0.07 ± 0.05
Melotte 66 −0.31 ± 0.07 0.01 ± 0.08 ... ... ... ... −0.29 ± 0.10 0.04 ± 0.09
Berkeley 32 −0.33 ± 0.17 −0.04 ± 0.14 ... ... ... ... −0.27 ± 0.10 0.02 ± 0.09
Collinder 261 −0.07 ± 0.09 −0.21 ± 0.07 ... ... ... ... −0.07 ± 0.02 −0.16 ± 0.05

Download table as:  ASCIITypeset image

3.1. Ce versus Y

In Figure 2 we plot [Ce/H] versus [Y/H] for all the stars for which both elements were measured. Filled and open circles indicate dwarf and giant stars, respectively. The figure shows that the data points for dwarfs are clustered around the 1:1 relationship and are thus consistent with a similar average enrichment in Y and Ce, within uncertainties; on the other hand, the points of the giants have systematically lower Ce abundances than Y. Interestingly, the same also holds for the Hyades where dwarf stars were analyzed and where indeed Y appears to be more abundant than Ce. In these stars, the 508.743 nm yttrium line gives [Y/H] values consistent with [Ce/H] and systematically smaller than those from the other yttrium lines. The discrepancy is larger than 0.3 dex in some stars, characterized by low temperature (below 4500 K), namely, 2, 6, 7 in Cr 261 and 1240 in Be 20, and again the [Y/H] from the 508.743 nm is consistent with [Ce/H] (see Table 4). Moreover, in all these stars the yttrium EWs are larger than 60 mÅ. In order to find some explanations for these observations, we considered the values [Y/Fe] derived from one yttrium line only (508.743 nm) for four OCs: M67, Be 32, Mel 66, and NGC 6253. These clusters have ages between 4 and 6 Gyr. Figure 3(a) shows our EW measurements of the men- tioned line versus the corresponding [Y/Fe] values. The points scatter around the center at [Y/Fe] ≈ 0. Performing the same exercise with the 520.041 nm yttrium line, a very different behavior appears. We show this in Figure 3(b). The M67 stars are centered on the [Y/Fe] ≈ 0 value, and indeed the EWs are smaller than 60 mÅ. The other points have larger EWs and are considerably shifted to the right side of the plot, hence to higher abundances, forming a plateau. For the 520.041 nm, the same happens for the remaining yttrium lines. The strict and systematic correlation between this shift to higher abundances and the EW values for which it happens (EWs larger than about 60 mÅ), together with the shape of the trend of Figure 3(b), are typical effects of saturation. Therefore, in the following we shall use Y abundances from the 508.743 nm line only for the evolved stars and for the Hyades members, because this line does not suffer from saturation for the EW values we measured.

Figure 2.

Figure 2. [Ce/H] vs. [Y/H]. Filled symbols denote dwarf stars, while open symbols indicate giant cluster members. The dashed line indicates the 1:1 relationship.

Standard image High-resolution image
Figure 3.

Figure 3. EW measurements vs. the corresponding abundances are plotted for two yttrium lines in four OCs: triangles refer to Be 32, squares to M67, crosses to Mel 66, and circles to NGC 6253. (a) Top panel: data from the 508.742 nm line. (b) Bottom panel: data from the 520.041 nm line.

Standard image High-resolution image

4. DISCUSSION

4.1. Evolution with Age

Figure 4, where we plot [X/Fe] ratios versus cluster age, summarizes our results for the evolution with age of the four elements analyzed in this study. The figure clearly shows that all four elements, and specifically their logarithmic ratios to Fe, increase with decreasing cluster age. All [El/Fe] ratios are consistent with the solar value, or are slightly below it, in old clusters. A steep growth is then seen between the ages of ∼1.5 and 0.5 Gyr, after which the [El/Fe] values remain constant around a plateau value of +0.2 dex. The distribution and growth of the first-peak elements are very similar to those of the second-peak ones. Also, although lanthanum and zirconium have a smaller number of points than yttrium and cerium because of the available wavelength range of the star spectra of each cluster, the measured [Zr/Fe] ratios are in very good agreement with [Y/Fe] in the clusters where both elements could be measured; the same happens for lanthanum and cerium. Our results confirm and strengthen those previously obtained by D'Orazi et al. (2009) for barium; indeed, our data are based on much higher statistics, as shown in Figure 5. This, in turn, makes clear that there is a need for an s-element production from very low mass AGB stars (with large ages, hence capable of affecting young clusters, not old ones) more effective than previously imagined. Together with the general trend of increasing s-element abundances in young clusters, some other features appear in the figure. First, we notice that at all ages a dispersion in s-process abundance is present. In particular, the Praesepe cluster lies significantly below the main trend and is less enriched than, e.g., the coeval Hyades cluster. On the other hand, three clusters with ages around 0.8–1 Gyr, NGC 2660, NGC 5822, and NGC 3680, are more enriched than other members of the sample with similar ages. We have no secure explanation for this dispersion. We suggest that it might not be intrinsic in part. Rather, it may be due to incorrect [Fe/H] determinations for these clusters. One example of this fact is probably shown by Praesepe, for which a high metallicity ([Fe/H] = 0.27) was measured (Pace et al. 2008). Its [El/H] values are in general agreement with those of other coeval clusters, but due to the high [Fe/H] estimate, the [El/Fe] ratios become much smaller. The high [Fe/H] value of Praesepe is actually rather strange for its Galactocentric radius, RGC. This peculiarity makes Praesepe a very interesting target for new measurements of the metallicity. On the other hand, part of the dispersion in s-process elements might instead be intrinsic and be related to the different positions in the disk of the sample clusters (see the labeled OCs of Figure 4 NGC 6192, IC 4651, NGC 6253, Cr 261 and the following paragraph).

Figure 4.

Figure 4. Average cluster [El/Fe] vs. cluster age. In the top panel, we show the evolution of first peak elements, namely, [Y/Fe] (circles) and [Zr/Fe] (triangles); in the bottom panel the second peak elements La (triangles) and Ce (circles) are shown. Error bars for Be 20 and Be 29 are not illustrated for graphical reasons.

Standard image High-resolution image
Figure 5.

Figure 5. Comparison of the evolutionary trends with cluster age for [Ce/Fe] and [Ba/Fe].

Standard image High-resolution image

We finally note the very high [Y/Fe] and [Ce/Fe] ratios of the old cluster Berkeley 29, which are both extremely above the trend. We will discuss the case of this cluster in the following section.

4.1.1. [El/H] versus RGC

The s-element enhancement can also be put in evidence by looking at the trend of [El/H] with RGC for clusters in different age intervals. This is shown in Figure 6, where the trends of [Y/H] and [Ce/H] with the Galactocentric radius are plotted. The figure suffers from low number statistics, in particular at large Galactocentric distances. Nevertheless it clearly shows that s-elements are characterized by a gradient along the Galactocentric radius similar to the one displayed by iron, iron-peak, and α elements (see Magrini et al. 2009). Most noticeably, two different gradients actually seem to be present. One is traced by older OCs (open circles of the figure), and one characterizes the younger objects (filled circles). The latter lies above the previous one, along the entire range of RGC (see Figure 7), suggesting that the increase in the s-element enrichment occurs across the whole Galactic disk, once very low mass stars with long evolutionary timescales have started to contribute.

Figure 6.

Figure 6. [El/H] vs. Galactocentric radius for yttrium and cerium. Filled and open circles represent clusters younger and older than 1.5 Gyr, respectively.

Standard image High-resolution image
Figure 7.

Figure 7. [Y/Fe] vs. Galactocentric radius. Filled and open circles represent clusters younger and older than 1.5 Gyr, respectively.

Standard image High-resolution image

Moreover, looking at Figure 7, the four labeled clusters with RGC smaller than ∼7.7 kpc show a different behavior with respect to the other OCs. These clusters, Cr 261, NGC 6253, IC 4651, and NGC 6192, are also labeled in Figure 4, where the trend they trace remains slightly below the main one. A possible explanation could be related to the infall process that occurred during the disk formation. Indeed, a rapid enrichment of iron in the inner zone of the disk, favored by mass infall from the halo, could shift the peak of star formation there toward higher [Fe/H] and [O/H]; the zones with larger RGC would be instead characterized by a star formation rate peaked at lower [Fe/H] values. Since a larger amount of iron and oxygen makes the occurrence of the third dredge-up in AGB stars more difficult and reduces neutron exposure (see Straniero et al. 2003 and references therein), it is reasonable that in the inner zone of the disk the s-element production was more limited. Further studies are required to extend the statistics of s-element measurements in the inner OCs (RGC < 7.5 kpc), while chemical evolutionary calculations at different RGC could verify the reliability of the hypothesis outlined above.

4.1.2. Berkeley 29

As mentioned above, both yttrium and cerium in this cluster show an anomalous enrichment. This result is not in agreement with the trend shown by other clusters, as is evident in Figure 4. Indeed, the age (4.3 Gyr) of this cluster should imply a solar [El/Fe] value. Members of this cluster are characterized by a dispersion in both yttrium and cerium, ranging from moderate to extreme enrichment.

In order to further check the high abundances by means of a differential comparison, we superimposed the spectra of two stars with similar parameters, namely, star 933 of Be 29 and star 938 of Be 32.

Figure 8 shows an example of this comparison for the yttrium line λ = 508.742 nm and supports the anomalous s-element enrichment found in Be 29. The signal-to-noise ratios of the two spectra (40–50 for 933, 50–60 for 938) do not allow us to be confident of the reliability of our EW measurements. Iron and α-element abundances of this cluster (see Sestito et al. 2008) have values that are compatible with the solar ones, so that our s-element abundances look strange. On the other hand, in favor of the enrichment there is the fact that all the analyzed lines of the four stars in Be 29 indicate an overproduction of yttrium and cerium. We note that Be 29 is also enriched in Ba (see Figure 5). Further investigations on s-element abundances in outer OCs are required and could become relevant clues for the knowledge of these external zones of the disk and their evolution.

Figure 8.

Figure 8. Yttrium line λ = 508.743 nm from the star 933 in Be 29 (solid line) and from the star 938 in Be 32 (dashed line). The comparison shows the larger EW of the line in the spectrum of 933.

Standard image High-resolution image

4.1.3. Clusters Studied in the Literature

Several recent studies focusing on abundance determinations in OCs include the measurement of one or more s-process elements. In Figure 9 we show the evolution of Y and Zr (upper panel) and Ce and La (bottom panel) obtained by using the most recent abundance determinations. Note that cluster ages are (or have been put by us) on the same scale as that used for the OCs discussed in this paper.

Figure 9.

Figure 9. Similar to Figure 4, but measurements from the literature are also shown. In the top panel we show the evolution of first peak elements, namely, [Y/Fe] (circles) and [Zr/Fe] (triangles); in the bottom panel the second peak elements La (triangles) and Ce (circles) are shown. Open and filled symbols denote abundances from this paper and the literature, respectively. Literature data include: Hamdani et al. (2000, NGC 2360); Tautvaisiene et al. (2005, NGC 7789); Carraro et al. (2008, NGC 2112); Jacobson et al. (2008, NGC 7142); Jacobson et al. (2009, NGC 1817, NGC 1883, NGC 2141, and NGC 2158); Friel et al. (2010, Be 31, Be 39, NGC 1193, and NGC 188); Pancino et al. (2010, NGC 7789, NGC 2420, Cr 110, and NGC 2209); Mikolaitis et al. (2010, NGC 6134); Jacobson et al. (2011b, NGC 2243).

Standard image High-resolution image

Measurements of s-process elements from the literature are highly non-homogeneous and difficult to compare with the present study because different lines, stars, and methods have been used. Nevertheless, the figure shows an overall fair agreement, as data from the literature follow the same trend displayed by our own measurements. The agreement is particularly good for Y, La, and Ce, while Zr is more scattered; specifically, a couple of relatively young clusters (NGC 1883 and NGC 7789) have a low [Zr/Fe] ratio, while a few very old ones (NGC 7142, NGC 1193, NGC 188, Be 31) are characterized by enhanced ratios. We note that for NGC 7789, enhanced [Y/Fe], [La/Fe], and [Ce/Fe] have been reported, so we believe that Zr was underestimated by Tautvaisiene et al. (2005). Zr determinations for NGC 1193, NGC 188, NGC 7142, and Be 31 come from Friel et al. (2010) and Jacobson et al. (2008). In both papers the abundances of Zr were based on at most two Zr i lines; large variations are seen within each cluster and, as stressed by Friel et al. (2010), abundances are highly uncertain. While new measurements of Zr and of the other s-process elements for all these clusters are needed, we conclude that the data from the literature generally confirm the main results of our paper.

4.2. Anomalously Enriched Star

Table 4 shows the presence of one anomalous star, which is significantly more s-element enriched than the other members of the same cluster. This star is 941 in Be 32. The discrepancy between its [El/H] values with respect to the mean of the cluster to which it belongs is evident. This enhancement is present both for Y and Ce, thus indicating that it is most likely real and not due to problems with the analysis.

In order to further investigate this point, we superimposed the spectra of this anomalous star to that of another star with similar stellar parameters and belonging to the same cluster, namely, star 19; this allows us to check if the discrepancy occurs because of possible errors in the determination of parameters like the star gravity or temperature. In Figure 10 we show an example of this exercise. The temperatures of the two objects are equal, while their log  g differs by 0.1 dex and their metallicities differ by 0.06 dex. The two spectra are perfectly in agreement with each other, with the exception of the Y ii line λ = 508.7 nm, which is stronger in the 941 spectrum. The agreement in the other regions of the spectra indicates that errors in stellar parameter determinations cannot be responsible for the anomalies in the s-element abundances found in 941. We performed this test for the other cerium lines, finding the same results as for yttrium. We thus conclude that the enrichment of this star is real.

Figure 10.

Figure 10. Comparison between the spectra of stars 941 and 19 in Be 32. The solid line represents star 941 and the dashed one star 19. The Y ii line is that at 508.743 nm. The upper plot shows the agreement between the two spectra, as a consequence of the similarity in the stellar parameters. The bottom panel is a zoom of the upper one and makes clear that there exists a discrepancy between the yttrium lines of the two stars.

Standard image High-resolution image

The above anomaly might be understood attributing the s-element enhancement to a mass transfer phenomenon in a binary system. If the anomalous star has a white dwarf companion, we might suppose that this was once the primary component of the system, with an intermediate mass (e.g., 3–4 M or more), and that it underwent s-processing during its AGB phase. Such a star would not have activated efficiently the 13C neutron source, but it would have nevertheless produced neutrons through the 22Ne(α, n) 25Mg reaction in the thermal pulses. This neutron source is actually the dominant one for all intermediate-mass (Lugaro et al. 2007) and high-mass stars (Busso & Gallino 1985; Raiteri et al. 1992). In this scenario, the secondary component we see today would be the equivalent of a Ba-star, i.e., it would have accreted s-process-rich material from the AGB primary.

5. CONCLUSIONS

  • 1.  
    Our analysis shows that the growth of barium in young OCs, originally found by D'Orazi et al. (2009), is accompanied by a similar increase of other neutron-capture elements belonging to both the first and second s-process peaks (yttrium, zirconium, lanthanum, and cerium). Specifically, we find that clusters older than about 1.5 Gyr have roughly solar [El/Fe] values, while younger clusters show an enhancement of about 0.2 dex. This growth terminates with a plateau for the youngest members of our sample, after which no further enrichment is seen. This suggests that the further increase in barium found by D'Orazi et al. (2009) should be attributed to other causes, affecting barium but not the abundance determinations of the other elements we have studied. NLTE effects, as mentioned by D'Orazi et al. (2009), could explain this behavior; a saturation of the barium lines they used is also a possibility.
  • 2.  
    The confirmation of the s-element enrichment in young OCs now needs a theoretical basis, including an enhanced production of neutron-capture nuclei by very low mass stars (M ⩽ 1.5 M), which are so far often neglected in dealing with s-processing in the Galactic disk. Very recently, models of extra-mixing phenomena that might be promising for the formation of an extended 13C pocket in very low mass stars have been discussed (Busso et al. 2007; Nordhaus et al. 2008). These are based on dynamo-driven magnetic buoyancy episodes in magnetically active stars (like those discussed in Andrews et al. 1988). These phenomena should also occur on the AGB (Busso et al. 2010; Palmerini et al. 2011) and would not require special conditions other than the presence of magnetized structures lighter than the environment. In particular, molecular weight inversions induced by nuclear burning are not necessary, contrary to the case with the commonly assumed mixing phenomena induced by thermohaline diffusion. Hence, they might also operate below the dredge-up in the interpulse phases of AGB stars, where nuclear sources are not present. Models of such mixing events and of the ensuing nucleosynthesis in LM-AGB stars should now be coupled to Galactic evolution models in order to ascertain whether the whole set of enhanced s-process abundances so far measured can be accounted for. We have already started new calculations with this specific scope.
  • 3.  
    We underline the high relevance for our analysis of the rather detailed knowledge achieved for OCs. We can derive their age through solid and homogeneous estimates. This was fundamental for obtaining our results. No clear trend would have been seen using only the usual [El/Fe] versus [Fe/H] analysis.
  • 4.  
    We identify one star that is anomalously s-process-rich. We interpret its anomaly as the consequence of a mass-transfer episode in a binary system, from a more massive, evolved companion that had already produced s-elements.
  • 5.  
    Finally, we underline that the models with enhanced s-element production, now required to explain our data, should be verified with the other test elements involved in the same He-shell burning phases where s-elements are produced. This is in particular the case of fluorine, whose abundance is enhanced in s-process-rich AGB stars (Abia et al. 2010). Observations also aimed at measuring the F abundance in MS and RGB stars of OCs should be carried out to allow for a more complete set of constraints to the upgrade of AGB nucleosynthesis models that is now clearly required.

We are grateful to C. Sneden and S. Andrievsky for their help during the abundance analysis. We also thank G. Pace for providing spectra of members of IC 4756, NGC 5822, Praesepe, M67, NGC 3680, and IC 4651. We acknowledge clarifying discussions on the s-process and on AGB nucleosynthesis with R. Gallino, O. Straniero, S. Cristallo, and C. Abia.

Footnotes

  • IRAF is distributed by the National Optical Astronomical Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation.

Please wait… references are loading.
10.1088/0004-637X/736/2/120